Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 May 1.
Published in final edited form as: Immunol Rev. 2020 Mar 12;295(1):167–186. doi: 10.1111/imr.12847

Metabolic determinants of lupus pathogenesis

X Teng 1, J Brown 1, SC Choi 1, W Li 1, L Morel 1
PMCID: PMC7180129  NIHMSID: NIHMS1574199  PMID: 32162304

Summary

The metabolism of healthy murine and more recently human immune cells has been investigated with an increasing amount of details. These studies have revealed that the challenges presented by immune cells to respond rapidly to a wide variety of triggers by adjusting the amount, type and utilization of the nutrients they import. A concept has emerged that cellular metabolic programs regulate the size of the immune response and the plasticity of its effector functions. This has generated a lot of enthusiasm with the prediction that cellular metabolism could be manipulated to either enhance or limit an immune response. In support of this hypothesis, studies in animal models as well as human subjects have shown that the dysregulation of the immune system in autoimmune diseases is associated with a skewing of the immunometabolic programs. These studies have been mostly conducted on autoimmune CD4+ T cells, with the metabolism of other immune cells in autoimmune settings still being understudied. Here we discuss systemic metabolism as well as cellular immunometabolism as novel tools to decipher fundamental mechanisms of autoimmunity. We review the contribution of each major metabolic pathway to autoimmune diseases, with a focus on systemic lupus erythematosus (SLE), with the relevant translational opportunities, existing or predicted from results obtained with healthy immune cells. Finally, we review how targeting metabolic programs may present novel therapeutic venues.

Keywords: Lupus, metabolism, glucose, glutamine, metabolic inhibitors

1. Introduction

The immune system has evolved to respond to challenges from pathogens with rapidly proliferating immune cells equipped with highly specialized effector functions that are performed changes in cellular programs that produce new effector molecules, such as antibodies and cytokines. The energetic requirements to meet these challenges have only been investigated recently, although immune cells, were routinely cultured in laboratories with large amounts of nutrients, especially glucose and glutamine. It was first shown in 2002 that the ligation of the co-stimulatory receptor CD28 upregulated glucose uptake by CD4+ T cells by activating the PI3K pathway1. This simple in vitro experiment was the first to establish a mechanism connecting cellular metabolism and immune activation. A requirement for glycolysis was established for B cell activation and differentiation a few years later, when it was shown that B cell receptor signaling also activated glucose uptake through the PI3K pathway2. These two seminal studies have launched the field of immunometabolism with an ever growing number of studies to characterize the energy needs of the immune system and to define the specific metabolic programs used by immune cells to achieve rapid growth and polarization into highly specialized effector functions3. Initially focused on the metabolism of immune cells, studies performed in mouse models or with patients with autoimmune diseases have later associated their impaired immune system with metabolic abnormalities, which were predicted to present novel therapeutic targets4,5. Early on, parallels have been established between rapidly proliferating activated lymphocytes and tumor cells both relying on glycolysis in the so-called Warburg effect6. Cellular metabolism research in oncology is still largely leading the field, with the goal of dissecting intricately dysregulated metabolic networks in both tumors and infiltrating immune cells7,8. Recent publications have reviewed how immunometabolism controls rheumatic diseases9, including systemic lupus erythematosus (SLE)10, which is, with rheumatoid arthritis, the autoimmune disease with the most advanced understanding of cellular metabolism.. This includes small clinical trials in which lupus patients have seen a reduction in disease activity following treatment with metabolic inhibitors11,12. Here, we review how the major metabolic pathways control the activation and differentiation of the immune cell types that are involved in lupus, as well as the studies that have been directly conducted in lupus pre-clinical models or in lupus patients. These studies are summarized in Table 1 along with the treatments with metabolic inhibitors that have been successful in pre-clinical models of the disease. Table 1 also includes metabolic inhibitors that should be evaluated in these pre-clinical models based on their efficacy on specific immune cell types that have a major impact on lupus development.

Table 1.

Metabolic drugs or treatment with reported efficacy in lupus patients or in mouse models of lupus, or in other immune cells with relevance to lupus

Drug / Treatment1 Metabolic target Immune target(s)2 SLE patients Lupus models
CR ? ? No effect (40) NZB/W (36, 37–39,41)
LXR agonist T0901317 Cholesterol efflux macrophages Pristane-induced diffuse alveolar hemorrhage (81)
metformin ETC complex 1, AMPK pDCs (120) B cells (139) CD4+ T cells (226) Reduced flares, prednisone-sparing (120) B6.Sle1.Sle2.Sle3 (121, 122) Roquinsan/san (123)
Rapamycin, sirolimus mTORC1 CD4+ T cells, effector memory CD8+ T cells (126) B cells (140) Reduced disease activity, prednisone-sparing (124, 125) Roquinsan/san (123) NZB/W (125)
Echinomycin, RNAi HIP1α Th1 and Th17(154, 155) GVHD (155) MRL/lpr (134)
CaMK4 inhibitor Glycolysis Th17 (165) Reduced Th17 in vitro (163) MRL/lpr (164, 166)
2DG Glycolysis CD4+ T cells (121, 122), Tfh cells B6.Sle1.Sle2.Sle3, NZB/W, B6.lpr, BXSB.Yaa, cGVHD (121, 122, 161)
PFKFB3 inhibitor glycolysis CD4+ T cells (169)
PKM2 stabilizer Glycolysis, STAT3-target genes Th1 and Th17 (171, 172)
Dimethyl fumarate glycolysis Th1 and Th17 (174) Discoid lupus (176, 177)
CG5, WZB117 Glut1-mediated glucose uptake CD4+ T cells (179) B6.Sle1.Sle2.Sle3, cGVHD (179)
NAC oxidation CD4+ T cells (208) Reduced disease activity and mTOR activation in CD4+ T cells (208)
MitoTempo mtROS Neutrophils, NETs (209) MRL/lpr (209)
4-octyl itaconate TCA cycle, Nrf2, Macrophages, Th17 (222, 223) Decreases inflammation in SLE PBMCs (223)
P2×7 agonist, BzATP Extracellular ATP Tfh and GC B cells (228) Pristane-induced lupus (228)
D-Mannose Decreases glycolysis, increases FAO Treg (233)
BCAT1 inhibitor Leucine transamination, TCA cycle macrophages Decreased inflammation in macrophages in vitro (241) Crescentic glomerulonephritis (242)
BPTES, DON glutaminolysis Th17, Tfh, GC B cells MRL/lpr (14) B6.Sle1.Sle2.Sle3, (161)
Soraphen A FA synthesis Th17 (259)
glucosylceramide synthase inhibitor glucosylceramide synthase CD4+ T cells (262) restores BTLA functionality in lupus CD4+ T cells (262)
1.

Drugs/ Treatments are listed in the order they appear in the text.

2.

Reported immune cells in which the drug has shown an effect. The list does not preclude other cell types that may also be targeted directly

2. Lupus and systemic metabolism

Immunometabolism commonly refers to the metabolic processes that regulate immune cell responses at the cellular level. Systemic nutrient availability and utilization by the whole organism, and how it may impact the metabolism of immune cells has not been considered to the same extent. It is however well established that obesity induces inflammation, and a recently published study showed that high dietary glucose promotes the differentiation of Th17 cells in mouse models of colitis and experimental autoimmune encephalomyelitis (EAE)13, a model of multiple sclerosis, an autoimmune disease with pathogenic T cells as SLE. Surprisingly, high dietary glucose did not increased circulating blood sugar nor enhanced glycolysis in T cells, a process necessary for Th17 polarization14. Instead, high dietary glucose accelerated TGFβ processing by increasing ROS production. Th17 cells have been implicated in lupus pathogenesis15. It is therefore tantalizing to predict that high glucose consumption may aggravate disease in lupus patients, a hypothesis that should be tested in mouse models. It should be noted it is not known whether a reduction in glucose consumption has a protective effect in autoimmunity.

Metabolic syndrome (MetS) corresponds to a cluster of metabolic abnormalities including hypertension, abdominal obesity, dyslipidemia, hyperglycemia and insulin resistance, in which chronic inflammation and oxidative stress have been proposed as potential mechanisms. SLE patients present a higher risk for developing MetS than healthy controls (HCs)16,17. A multivariate analysis has associated MetS with disease activity and steroid exposure in a cohort of SLE patients18. Since MetS increases the risk of cardiovascular events in the general population19, it is also likely that SLE patients with MetS are also at higher risk for cardiovascular disease. Indeed, MetS has been associated with organ damage and vascular events20, as well as increased subclinical atherosclerosis in SLE patients21. Obese women followed over a period of 20 years as part of the Nurses Health Studies (NHSII) had an increased risk of developing SLE compared to lean controls22, and obesity during teenage years increased this risk further23. As another part of manifestations commonly associated with MetS, SLE patients have an increased frequency of dyslipidemia24. Interestingly, dyslipidemia expands the number of autoimmune follicular helper T (Tfh) cells and germinal centers (GC)25. It is therefore possible that dyslipidemia contributes to the expansion of the number of Tfh cells in SLE patients, a phenotype that correlates with disease activity26. Therefore, these results obtained in a mouse model suggested that dyslipidemia enhances autoimmune pathology. Collectively, these findings suggest that MetS could contribute to the development of cardiovascular complications in SLE patients. The validation of a mutual enhancement between MetS and SLE will requires detailed epidemiologic studies in large cohorts of patients as well as mechanistic studies in mouse models combining features of both diseases.

One of such mechanistic studies was conducted in ApoE or Ldlr-deficient chimeric mice carrying the bone marrow of lupus-prone mice and fed with an atherogenic Western diet. In these mice, dyslipidemia increased the production of autoantibodies and the severity of autoimmune pathology27. The mechanism responsible for these findings was identified as an expansion of the number of Tfh cells by IL-27 produced by DCs in response to lipids in a TLR4-dependent manner27. The importance of this pathway was validated by blocking IL-27, which reduced autoantibody levels and the number of Tfh cells in the chimeric atherogenic mice. Furthermore, higher IL-27 concentrations were found in patients with hypercholesterolemia. An independent study has reported elevated levels of IL-27 produced by DCs in lupus-prone mice, as well as in SLE patients, which were directly dependent on type I IFN signaling28. These results suggest a link between IL-27, type I IFN, autoimmunity and dyslipidemia, with a mutual enhancement of the two types of pathogenesis. ApoA1-deficiency greatly reduces HDL levels in mice, which present cholesterol-engorged enlarged lymph nodes and also develop an autoimmune syndrome with anti-dsDNA autoantibodies and a spontaneous activation of CD4+ T cells29. These reverse genetic experiments suggested that the dyslipidemia induced by an impaired cholesterol import or efflux promotes systemic autoimmunity.

On the opposite spectrum from MetS, calorie restriction (CR) has been shown to modulate immune cells toward less inflammatory phenotypes, some of which may be directly involved in lupus pathogenesis. T cell phenotypes associated with aging, including memory cell accumulation and reduced naive cell frequencies, were both attenuated by CR30. Another study has found however that memory T cells accumulated in the bone marrow during CR where they were protected and displayed enhanced functionality31. Additionally, fasting reduced the number of gut-associated B cells in Peyer’s patches, directly by apoptosis for GC B cells and indirectly by migration to the bone marrow for naïve B cells32. Upon re-introduction of calories, naïve B cells migrated back to Peyer’s patches32. Therefore, it is possible that CR may modulate B cell populations and autoantibody levels in lupus. Within the innate immune system, circulating monocyte frequency and functionality were reduced in response to fasting in both humans and mice, which improved clinical parameters in EAE33. Similarly, intermittent fasting was protective in EAE34,35, and protection was mediated at least in part by microbiota changes35. Protective microbiota changes have also been observed in response to CR in a model of induced colitis and in human inflammatory bowel disease36, suggesting that it may be a major mechanism for the immunomodulatory effect of CR.

Evidence suggests that CR is protective in murine models of SLE37 (Table 1). CR in NZB/W lupus-prone mice prolonged lifespan38, prevented renal immune complex deposition39, and modulated renal aquaporin expression40, the latter of which is associated with renal disease41. CR also suppressed lupus nephritis in NZB/W mice in parallel with a reduced presence of platelet-derived growth factor (PDGF) and thrombin receptor in renal tissue, which are known to be involved in glomerular inflammation42. These studies suggest that CR attenuates parameters associated with kidney pathology in murine models of lupus. Remarkably, CR delays and attenuates disease in NZB/W mice to a similar extent as cyclophosphamide, a cytotoxic drug that is used in the clinic to treat severe lupus, by reducing autoantibody levels, B cell activation, and IL-10 production in addition to increasing IL-2 levels37. Additionally, CR decreased pro-inflammatory cytokine production in NZB/W mice43. Overall, studies suggest that CR has a protective effect in lupus by modulating autoimmune phenotypes in murine models. However, these promising results have not been validated in SLE patients, in which CR did not change disease activity44. This study was however limited by very short (6-weeks) duration and a small cohort size, and did not account for the use of immunomodulatory drugs. Therefore, the potential benefits of CR in SLE should be still considered as an open question. Moreover, the specific mechanisms of the potential beneficial effects of CR in lupus have not been elucidated. On the opposite side of systemic metabolism, it would be of great interest to investigate whether the increased frequency of Th17 cells observed in mice with a high glucose intake13 could also accelerates lupus pathology.

It is also possible that the effects of CR are indirect and mediated by adipokine alterations. Adipokines are a group of peptides including leptin and adiponectin synthesized by white adipose tissue (WAT) that regulate energy homeostasis, but are also involved in inflammation30. Altered levels of adipokines have been reported in patients with MetS45 and with SLE, as detailed below. Leptin regulates energy homeostasis in the neuroendocrine system by increasing satiety to simultaneously limit food intake and promote energy utilization. Leptin contributes to inflammation through both innate and adaptive mechanisms, which may be due to its structural similarity to several pro-inflammatory cytokines, including IL-646. High levels of circulating leptin correlate with disease activity in SLE patients21,4751. Fasting decreases the production of leptin and expands the frequency of FOXP3+ regulatory (Treg) cells in SLE patients leading to a reduced disease activity52. However, SLE patients with reduced serum leptin have also been reported in some studies53,54, while others show no difference between SLE patients and HCs55. Genetic variations in the leptin and leptin receptor genes are associated with some clinical parameters of SLE such as photosensitivity and pericarditis, in spite of a lack of association with genetic susceptibility to SLE56. High leptin levels in SLE patients correlate with vascular stiffness parameters, increased risk for atherosclerosis, and with increased inflammatory atherosclerotic biomarkers57,58. Furthermore, high levels of circulating leptin correlate with subclinical atherosclerosis in SLE patients with MetS21. Additionally SLE patients with plaque present higher leptin levels than SLE patients without plaque58. Therefore, high leptin levels may link some of the manifestations of MetS and SLE.

Leptin levels also influence immune cell functions. Leptin signaling is necessary for murine Th17 effector functions59, and it skews human and mouse T cells toward a Th1 phenotype in vitro60,61, which is relevant to SLE patients who present an increased frequency of Th1 cells and IFNγ production. Transgenic expression of leptin in pigs results in a lupus-like disease characterized by increased anti-dsDNA antibodies, renal immune complex deposition, nephritis, and decreased frequency of Treg cells62. Leptin levels inversely correlate with the frequency of Treg cells in SLE patients48. Leptin deficiency in MRL/lpr lupus-prone mice was protective for disease development and it increased the frequency of Treg cells63. Likewise, leptin inhibited Treg cell polarization in vitro and its neutralization resulted in Treg cell expansion in NZB/W mice64. In this same study, leptin administration accelerated disease, whereas blocking leptin delayed disease progression64. Leptin deficiency in MRL/lpr mice also lowered serum IL-17 levels63. Similarly, leptin enhanced Th17 responses in NZB/W mice by inducing RORγt expression63,65 and increasing Th17 polarization in vitro63. Leptin has also been shown to increase glycolysis in murine T cells and consequently enhance effector function66. Therefore, there is a consistent association between high leptin levels and enhanced pro-inflammatory effector T cell populations. In addition to its effects on T cells, leptin activates B cells and augments the production of pro-inflammatory cytokines as well as that of IL-1067,68. Interestingly, these T cell phenotypes may be induced indirectly by leptin promoting the clearance of apoptotic cells by macrophages, which increases the availability of phagocytized self-antigens to stimulate autoreactive T cells in NZB/W mice69,70. In summary, leptin activates B cells, increases the frequency of Th17 and Th1 cells while decreasing the frequency of Treg cells, and it may due at least in part to increased self-antigen availability, all of which are lupus-relevant phenotypes.

Adiponectin is another adipokine synthesized by WAT and it is generally considered to be an anti-inflammatory cytokine. Adiponectin gene polymorphisms are not associated with SLE risk71, but high levels of adiponectin have been reported in SLE patients with and without lupus nephritis. Correlations with disease activity are however contradictory50,55,7274. PBMCs of SLE patients present higher amounts of adiponectin transcript and protein as compared to HCs, but there is no correlation between adiponectin expression levels and clinical parameters or disease activity indexes75. Moreover, increased adiponectin levels in MRL/lpr mice do not promote disease progression76. Thus, both adiponectin and leptin levels are increased in lupus, but only leptin appears to contribute to disease phenotypes by modulating immune cell functions.

3. Cholesterol homeostasis in lupus

The function of immune cells is regulated by cholesterol through multiple mechanisms. First, cholesterol is an integral part of all cell plasma membranes, and a limiting factor for membrane synthesis, and hence, for cellular proliferation. A second function of cholesterol that is more specific to immune cells is its regulation of lipid raft assembly, thus regulating signaling in major pathways such as TCR, BCR TLRs and MHC, which are all lipid-raft dependent7779. A direct demonstration of the immune regulatory role of cholesterol through this pathway was provided by the analysis of mice with ApoE-deficient dendritic cells (DCs), in which the impaired removal of cholesterol from the membrane led to an accumulation of lipid rafts. This promoted MHC-II clustering and increased antigen presentation, which, in turn, expanded pro-inflammatory CD4+ T cells that supported skin allograft rejection, independently of dyslipidemia80

Cholesterol homeostasis balances synthesis, import into the cells, and efflux outside the cell. Cholesterol is synthesized by the HMG-CoA reductase, the therapeutic target of statins, and whose expression is regulated by the transcription factor SREBP81. Cellular import of cholesterol is performed by apolipoprotein E (ApoE) and the low density lipoprotein receptor (LDLR). The opposite efflux of cholesterol from immune cells is performed by apolipoprotein A1 (ApoA1) and the high density lipoprotein receptor (HDLR) mediate through the expression of liver X receptor (LXR)-regulated genes, such as the ATP-binding cassette transporters ABCA1 and ABCG1.

Cholesterol synthesis is low in immune cells relative to other cell types such as hepatocytes. Directly relevant to SLE, type I IFN promotes cholesterol import over synthesis82. This observation predicts that lupus-immune cells that are exposed to high levels of type I IFN would depend on a high cholesterol uptake for their proliferation and activation. Therefore, LXR agonists that promote cholesterol efflux should have a beneficial effect on lupus pathology. This hypothesis was verified in a pristane-induced model of diffuse alveolar hemorrhage, in which the protective effect of an LXR agonist reprogrammed IFN-induced M1 macrophages to an M2-macrophage like phenotype83. Similarly, LXR-deficiency resulted in the development of systemic autoimmunity in mice with a defect in apoptotic cell clearance, while a LXR-agonist reduced autoimmune manifestations in the B6.lpr model of lupus84. The contribution of LXR-regulated cholesterol homeostasis was also demonstrated by the analysis of Abca1/g1 deficient mice, which presented lymphadenopathy and glomerulonephritis85,86. Interestingly, DCs were the functional cell type in these mice in which alterations in cholesterol efflux was responsible for immune activation. Abca1/g1 deficiency activated the inflammasome in DCs, which proliferated and secreted cytokines that ultimately expanded lupus-associated effector lymphocyte populations, namely Th1, Th17 and Tfh cells, GC B cells, and plasma cells86. As expected, Abca1/g1 deficiency enhanced lipid-raft associated signaling, such as inflammatory responses to TLR4 activation in macrophages87, and TCR signaling and proliferation in T cells88. It is not clear, however, why, contrary to DCs, these phenotypes did not lead to autoimmunity, and further studies are necessary to better understand cell-specific and intrinsic mechanisms of immune cell activation by cholesterol efflux. Given the association of MetS and lupus, it would also be of great interest to link circulating cholesterol levels with cellular immune metabolism and activation.

4. Metabolite profiles in lupus

Metabolite profiling of SLE patients has revealed differences with HCs8992. Serum metabolite profiles also differ between patients separated by disease activity scores89, and between SLE patients with and without lupus nephritis92. Additionally, the metabolome of SLE patients is different from that of patients with other autoimmune diseases such as primary Sjogren’s syndrome and systemic sclerosis89, suggesting that these metabolic signatures are not a global result of autoimmunity, but rather are specific to SLE. Compared to HCs, SLE patients have increased metabolites related to oxidative stress8991, and decreased amino acids8992. Some decreases of amino acids such as arginine89 could increase the levels of nitric oxide metabolites related to oxidative stress, which play a role in the pathology of SLE93. One of the amino acids with the most altered level in SLE patients is tryptophan89,9496. SLE patients have depleted levels of tryptophan, and opposite alterations in the levels of its proximal metabolites, decreased serotonin and increased kynurenine levels9499. It is possible that these metabolite alterations are a reflection of upregulated indolamine 2,3-dioxygenase (IDO1) expression94,95. In fact, it has been hypothesized that an increased IDO1 expression in lupus is due to the type I IFN signature observed in patients95, as IFNα is known to upregulate IDO1. IDO1 protein levels were increased upon in vitro stimulation of a cell line with serum from SLE patients95. Regardless of the mechanisms responsible for the altered tryptophan metabolism, it is correlated with disease activity in SLE patients. Depleted serotonin in SLE patients inversely correlated with nephritis and anti-dsDNA autoantibodies95 and the kynurenine/tryptophan ratio correlated with severe fatigue94. Kynurenine was also one of the most increased metabolites in SLE peripheral blood leukocytes (PBLs) and it allowed for discrimination between responder and non-responder SLE patients to a treatment with N-acetylcysteine (NAC)100. Furthermore, kynurenine activated mTOR in healthy PBLs100, suggesting that this metabolite may directly influence cellular metabolism. Along with altered amino acids, dyslipidemia and hypercholesterolemia were also found in SLE patient metabolite profiles8992,101103. 25-hydroxyvitamin D3 is significantly depleted in SLE serum104106, suggesting that the loss of immunomodulatory vitamin D metabolites may play a role in pathogenesis. Indeed, vitamin D deficiency inversely correlates with SLE disease activity in addition to other clinical parameters105,107.

A recent analysis found significant differences in fecal metabolite profiles between SLE patients and HCs108, which ultimately could allow for a non-invasive screening for SLE-associated biomarkers. Similar to serum profiles, the most altered fecal metabolites were related to amino acid, lipid, and vitamin metabolism, as well as an enriched nitrogen and tRNA biosynthesis in SLE patient fecal contents108. Fecal metabolomics may be important considering increasing reports of altered SLE fecal microbiomes both in patients and murine models of lupus109115. Pathway analyses of serum metabolite profiles also suggested an altered gut microbial metabolism in SLE patients compared to HCs101,116, including an increase in total free fatty acids (FFAs) and short chain fatty acids (SCFAs) synthesized by bacteria112. Additionally, disrupted metabolic pathways were found in the urine of SLE patients compared to HCs including amino acid, TCA cycle, and purine/pyrimidine metabolism116.

A major challenge in metabolomic studies in SLE is the use of cytotoxic and immunosuppressive drugs, such as steroids, which could themselves shift metabolite profiles. In addition, the links between systemic and cellular metabolic alterations are largely unknown, and it is not clear if alterations in systemic levels of specific nutrients impact availability at the cellular level. Nonetheless, studies have highlighted energy homeostasis, oxidative stress, and amino acid metabolism as major global metabolic pathways that are disrupted in SLE patients.

5. MTOR activation in lupus

Mammalian target of rapamycin (mTOR) and AMP-activated protein kinase (AMPK) are major sensors of the cellular energy status, which represents a major regulation of the activation and differentiation of immune cells117. High levels of AMP, in response to either poor nutrient availability and reduced ATP production, or high ATP consumption activate AMPK. Activated AMPK restores intracellular ATP levels by reducing its consumption, mostly by inhibiting protein synthesis, while promoting ATP synthesis through metabolic processes such as fatty acid oxidation (FAO) and glucose uptake, as well as cellular processes such as autophagy118. Activated AMPK also inhibits the activation of mTOR, a kinase complex that promotes energy consuming processes. The catalytic subunit of mTOR signals through two multimeric complexes, mTORC1 and mTORC2, which only differ by their scaffold proteins RAPTOR and RICTOR, respectively. Activation of both complexes promotes glucose metabolism, linking mTORC and glycolysis.

Metformin is a drug with multiple targets, one of which is the activation of AMPK119. SLE patients that received metformin as an add-on treatment to their standard of care showed disease improvement120. Metformin also showed some efficacy in mouse models of lupus121123. mTOR activation has been to be directly implicated in the CD4+ T cell phenotypes of lupus-prone mice121 and lupus patients124. It has been recognized early that treatment with rapamycin, an mTOR inhibitor, reduced the severity of disease in nephritic NZB/W mice125. More importantly, the therapeutic efficacy of inhibition mTOR has been demonstrated in patients with refractory SLE, in which treatment with sirolimus improved disease activity and reduced prednisone exposure126, two major end-points in lupus clinical trials. Sirolimus may also have therapeutic potentials for patients with refractory lupus nephritis127. The encouraging results obtained by these two independent one-arm, open-label, retrospective studies indicate a need for follow-up double-blind larger clinical trials in ethnically diverse patient populations to identify the lupus patients that may benefit from mTOR blockade. Further, the identity of the effective cellular targets of AMPK activation by metformin or mTOR inhibition in lupus patients is unclear. Surprisingly, a reduced number of effector memory CD8+ T cells was the best predictor of the therapeutic response to sirolimus in SLE patients126, although the role of this T cell subset is yet undefined in lupus. These findings urge caution in the extrapolation of studies with in vitro treatment of purified immune cell populations or cell-specific genetic targeting, which may lead to significant results, but with little relevance to the action of the drug when administered to patients.

Studies in mice have shown that mTOR activation promotes the differentiation of Th1, Th17128, as well as Tfh T cells129, which are T cell subsets associated with lupus130,131. On the other hand, AMPK activation promotes the expansion of the Treg cell subset. Treg cells in which AMPK activation is impaired by the deletion of Lkb1, its upstream kinase, presented an impaired function that promoted a Th2-dominant severe autoimmune phenotype132,133. A complex role of mTOR in Treg cells has emerged from genetic targeting experiments. MTOR limits the maintenance of long-lived central Treg cells, but promotes the differentiation of effector Treg cells134. Mtor deficiency reduced the frequency of Treg cells, resulting in the expected corresponding spontaneous effector T-cell activation that has been reported for other Treg altering mutations135. PP2A is a serine-threonine phosphatase that prevents the development SLE by limiting the production of IL-2 and IL-17 by CD4+ T cells136. Pp2a-deficient Treg cells were impaired as a result of an increased mTORC1 activation, resulting in a severe systemic autoimmune phenotype137. Overall, these results showed that Treg cells require an intermediate level of mTOR activation, and that dephosphorylation by PP2A is one of the mechanisms by which this balanced activation is maintained. Follicular regulatory T (Tfr) cells are GC-specific Treg cells, and the number or function of Tfr cells may be defective in lupus patients138. As for their Treg cell precursors, the function and differentiation of Tfr cells also requires mTORC1-activation139. Overall, these results suggest a profound mTOR-dependence of the T cell subsets that have been associated with lupus and that limiting mTOR activation in CD4+ T cells is required to prevent the development of systemic autoimmunity.

In addition to CD4+ T cells, mTOR is over-activated in the B cells of SLE patients, and it correlates with plasmablast numbers and disease activity140. In the Roquinsan/san lupus-prone mouse, treatment with either metformin or rapamycin inhibited B cell differentiation into GC B and plasma cells, and reduced disease activity, which implicated the AMPK/mTOR pathway in the activation of autoreactive B cells123. These results also suggest that a high basal level of mTOR activation may set a lower threshold for B cell activation and differentiation, although neither study in SLE patients or in the Roquinsan/san mice could distinguish B-cell intrinsic mTOR activation. In fact, the role of mTOR in B cells is not well understood, with evidence for contributions to B cell development, differentiation, survival and function that may differ between specific differentiation stages141. Genetic ablation showed that a more preeminent role of mTORC1 activation in the early pro-B to small pre-B cell stages in the bone marrow as compared to the peripheral resting immature and recirculating mature B cells142. Negative selection of self-reactivity is a critical checkpoint for transitional B cells, which then mature into resting follicular B cells that achieve metabolic quiescence by simultaneously dimming mTORC1 and activating AMPK in both humans and mice143. In peripheral B cells, deletion or inhibition of the mTORC1 pathway reduced GC development, the production of high-affinity antibody, and class-switch recombination in mice after immunization144,145. Neither inducible deletion of Raptor nor acute treatment with rapamycin had measurable effects on either naive or antigen-specific memory B cells after immunization146, suggesting that mTORC1 activation is not required for these B cell subsets. However, both rapamycin and Raptor deletion in B cells eliminated newly formed plasma cells and pre-existing GCs146. Interestingly, antibody production was decreased in a reversible manner by these two approaches, but the frequency of long-lived bone marrow plasma cells or their survival were unaffected146. Opposite results were obtained in plasma cells in which Tsc1, a negative regulator of mTORC1, was deleted. Tsc1−/− plasma cells showed an enhanced antibody production and its associated endoplasmic reticulum stress response, but displayed shorter lifespans147. The same phenotypes were displayed by plasma cells deficient in Atg5148, a master regulator of autophagy, a process that is inhibited by mTORC1. These results suggest that long-lived plasma cells may require mTOR activation to maintain longevity, but that the production of autoantibodies by short-lived plasma cells may be enhanced by mTOR inhibition. Lupus autoantibodies are produced by both types of plasma cells149, which illustrates the potential complexities of metabolic regulation of immune cell function in the setting of complex autoimmune diseases such as SLE. The effect of mTOR inhibition by sirolimus in SLE patients was not determined in B cells, and only a modest reduction of some autoantibodies was observed126, although the study was underpowered for the analysis of treatment outcomes on autoantibody production. It is therefore still unclear whether lupus B cells or plasma cells have a functionally over active mTOR. Treatment of mice with Torin1, which inhibits both mTORC1 and mTORC2 signaling150, reduced modestly the numbers of immature, marginal zone and transitional B cells146. These results compared to the effect of rapamycin144146, which targets primarily mTORC1, may suggest a role for mTORC2 signaling in naive B cell development and homeostasis, while mTORC1 affects the later stages in development151. Overall, more studies are needed on the consequences of either mTOR deletion or over-activation in the B cell subsets that are directly relevant to lupus, such as transitional B cells, GC B cells, short- and long-lived plasma cells. It would be also of great interest to determine whether that a high basal level of mTOR activation may set a different tolerance threshold allowing the escape of autoreactive B cells.

The role of mTOR in DCs is largely unknown, but it has been proposed that it integrates pattern recognition signals with energy status for optimal DC activation and effector functions152. DCs, including plasmacytoid dendritic cells (pDCs), are active participants in lupus pathogenesis, but it is unknown whether it involves mTOR activation. Overall, multiple studies with human cells and in mouse models have converged to demonstrate a central role for mTOR in lupus by potentially affecting multiple cell types. The therapeutic efficacy of the modulation of the mTOR – APMK axis with sirolimus and metformin, respectively, provides an incentive for detailed mechanistic studies in mouse models as well as in lupus patients.

6. The HIF pathway in lupus

Hypoxia-induced transcription factor HIF1α, the master regulator of the cellular response to hypoxia, exerts potent effects on the immune system that are not directly related to oxygen levels, such as the activation of glycolysis with the associated pro-inflammatory consequences described in the next section153. Th17 polarization is HIF1α-dependent154, which suggested that targeting HIF1α could be beneficial in lupus. Th1 and Th17 differentiation was reduced by inhibiting HIF1α with echinomycin, and treatment with this drug reduced the manifestations of acute graft-versus-host disease (aGHVD) in a mouse model155. Furthermore, inhibition of Hif1α with RNAi ameliorated disease in MRL/lpr lupus-prone mice, which presented a strong reduction of IL-17 production134. Interestingly, HIF1α protein expression in Th17 cells is regulated by glutaminase GLS1, through which glutaminolysis indirectly controls glycolysis in this T cell subset15. Accordingly, MRL/lpr mice treated with the GLS1 inhibitor BPTES showed attenuated lupus outcomes in a Th17-dependent manner, and BPTES decreased the polarization of Th17 cells in SLE patients15.

Specific tissue microenvironments such as tumors or the inflamed kidneys in lupus nephritis156 are hypoxic. Directly relevant to humoral autoimmunity and lupus, it has been recently recognized that cells in the GC microenvironment have to cope with low amounts of oxygen. A dynamic balance between metabolic activation and inhibition occurs in GC B cells in response to oxygen sensing157,158. Expression of HIF1α is higher in GC B cells than in other splenic B cells. In the GC light zone (LZ), hypoxia promotes a higher glycolytic rate, which increases B cell apoptosis, diminishes proliferation and impairs immunoglobulin class switching by limiting AID expression157. These features are required for the antigen-driven selection process in the GC LZ. Sustained hypoxia or HIF stabilization inhibits mTORC1 activity in B lymphoblasts in the DZ, which impairs their proliferation and class-switching157. These findings demonstrate that oxygen sensing and rapid switch to the corresponding metabolic program is an essential requirement of GC B cells. Cell-specific deletions as well as in vitro activation in hypoxic conditions have also shown that Hif1a expression is necessary for optimal CD4+ T cell effector functions, including cytokine secretion and co-stimulation of B cells in humoral responses159. The production of high affinity autoantibodies by GC B cells that receive T cell help is highly relevant to lupus, and whether lupus GCs are regulated by the HIF pathway in the same manner as what has been described for non-autoimmune GCs is a question that remains to be answered.

7. Glycolysis and Lupus

Glycolysis refers to the conversion of glucose to pyruvate through eight metabolic intermediates in a process that generates two molecules of ATP. It also produced NADH, which donates an electron to complex I of the electron transport chain (ETC) to initiate oxidative phosphorylation (OXPHOS). Pyruvate has the one of two fates: oxidation in the Krebs cycle to produce of up to 38 molecules of ATP per molecule, or reduction into lactate when rapid cellular proliferation boosts the demand for metabolite intermediates, such as NAD. Glycolysis commonly refers to the lactate reduction of pyruvate, as opposed to glucose oxidative or mitochondrial (mt) metabolism.

Lupus CD4+ T cells in either SLE patients or spontaneous mouse models are characterized by a high level of glucose metabolism121, which is largely oxidation in the mt122. The high level of oxygen consumption and oxidation that have been reported in murine and human lupus CD4+ T cells121,160 is most likely largely used for this process. Glucose uptake through glucose transporters represents the first rate-limiting step of glycolysis. Glucose transporters are members of the large family of solute transporters (Slc) that serve as gate keepers of nutrient uptake and trafficking between cellular compartments. Only a fraction of Slc members have well characterized substrates and functions. Besides Glut1, the role of the other glucose transporters in immunometabolism is unknown, even when a differential expression has been validated, such as for Glut6, which is overexpressed in the spontaneous Tfh cells of lupus mice161. Glut1 is the glucose transporters that is expressed at the highest level by T cells upon TCR and CD28 signaling162. Mice transgenic for Glut1 accumulate activated CD4+ T cells, produce autoantibodies and present an immune complex deposition in the glomeruli akin to an early stage lupus nephritis163. Furthermore, Glut1 overexpression increased the numbers of Tfh and GC B cells, leading to elevated IL-21 and IgA production129. Although overexpression of Glut1 in mice was not sufficient to induce lupus, SLE patients with active disease present a higher GLUT1 on effector memory CD4+ T cells as compared to either HCs or patients with inactive disease, and this elevated GLUT1 level was associated with an elevated expression of calcium/calmodulin-dependent protein kinase 4 (CaMK4)164. CaMK4 is a major regulator of pathogenic T cells in lupus through its regulation of IL-2 and IL-17 expression165. Accordingly, the polarization of human Th17 cells, an effector subset that is highly dependent of glycolysis154, was reduced by CaMK4 inhibition, which reduced GLUT1 expression. Moreover, CaMK4 inhibition decreased the production of glycolytic intermediates by activated CD4+ T cells from MRL/lpr lupus prone mice164. These results showed a functional link between CaMK4 and glycolysis in T cells and suggests that the beneficial effect of CaMK4 inhibition in lupus166 may be due at least in part through the mitigation of immunometabolism.

Importantly, these studies have translational implications by predicting that the inhibition of glycolysis would have beneficial effects in lupus. Treatment of lupus-prone mice with 2-Deoxy-D-glucose (2DG), a glucose analog that inhibits the first reaction of glycolysis, prevented disease, but had only partial therapeutic effects121,122. 2DG was sufficient to eliminate the production of autoantibodies, but had little effect on the IFNγ-producing effector memory CD4+ T cells161. The combination of 2DG with metformin, which inhibits complex I of the mt electron transport chain (ETC)167, reduced the number of these pathogenic T cells and reversed lupus pathogenesis in mice121. These results indicate that targeting glycolysis, either alone or in combination with other drugs such as metformin, presents a therapeutic potential for lupus12. 2DG has been used in clinical trials in oncology without successful outcomes and with some safety concerns168. Other glycolytic enzymes have been targeted in mouse models in which inflammatory and/or autoreactive T cells play a role. Inhibiting 6-phosphofructo-2-kinase/fructose-2,6-biphosphatase 3 (PFKFB3) was very effective in controlling alloantigen activation of CD4+ T cells in a GVHD model169. Pyruvate kinase M2 is a glycolytic enzyme that also activates STAT3-mediated transcription when it translocates as a dimer to the nucleus. Blocking this process by enforcing the tetramer configuration reduced glycolysis and the production of inflammatory cytokines in macrophages obtained from patients with atherosclerosis170. Perhaps more relevant to lupus, PKM2 is required for Th1 and Th17 polarization171. Accordingly, preventing PKM2 dimerization also reduced glycolysis in CD4+ T cells and was protective in EAE171,172. Inhibition of either PFKFB3 or PKM2 has not been tested yet in lupus models. Pyruvate dehydrogenase (PDH) represents an essential node in glucose utilization by directing pyruvate to OXPHOS rather than glycolysis. PDH is converted from an inactive to an active form by pyruvate dehydrogenase phosphatase catalytic subunit 2 (PDP2), which is directly involved in Th17 polarization in lupus. Indeed, memory Th17 cells from SLE patients present a reduced PDP2 expression and over-expression of PDP2 in CD4+ T cells from either MRL/lpr mice or SLE patients suppressed Th17 differentiation. This PDH checkpoint of energy production may be ultimately regulated by the transcription factor ICER/CREM in Th17 cells173, in a model consistent with the glycolytic requirements of Th17 cells154 and the expansion of Th17 cells in SLE patients130. Finally, dimethyl fumarate, a derivative of the Krebs cycle intermediate fumarate, inactivates GAPDH, which is upstream of pyruvate, therefore inhibiting both branches of glycolysis. Dimethyl fumarate decreasing the differentiation and function of Th1 and Th17 cells, and ameliorates disease in EAE174. Dimethyl fumarate is effective in treating multiple sclerosis175, and discoid lupus in small open-label clinical trials176,177, in which the cellular and molecular targets of the drug have not been investigated.

As an alternative to inhibit glycolytic enzymes, the pharmacological targeting of GLUT1 has been considered to dampen the pathology induced by autoreactive T cells178. A proof-of-principle of the efficacy of this approach was obtained with a compound called CG-5, which blocked glucose uptake and glycolysis in murine and human CD4+ T cells179. CG5 ameliorated autoimmune phenotypes and reduced the production of autoantibodies in a spontaneous and an induced model of lupus. In these lupus models, CG-5 reduced the numbers of activated CD4+ T cells and GC B cells, inhibited Th1 and Th17 differentiation and also promoted Treg cell induction.

Multiple studies have shown that effector T cells have a higher requirement for glycolysis than naïve resting T cells, but this requirement varies among effector subsets, with Th17 and Th1 cells being highly glycolytic180. These studies have been however largely performed with in vitro polarization conditions, and a sophisticated isotope tracing analysis has challenged the high glycolysis requirement of activated T cells in physiological conditions in live mice181. Tfh cells in lupus mice present a high level of mTORC1 activation, and they are the most glycolytic CD4+ T cells in these mice. Accordingly, the frequency of Tfh cells, as well as that of GC B cells and the resulting production of autoantibodies, were reduced to non-autoimmune levels by a treatment with 2DG in several mouse models of lupus161. Tfh cells from lupus mice express high levels of the lactate transporter Mct4 and produce high amounts of lactate, measured as extracellular acidification rate161. This suggests that autoreactive Tfh cells require glucose for anaerobic glycolysis. It has not been excluded, however, that mt metabolism of glucose is also required for the function of autoreactive Tfh cells.

The inhibition of glycolysis also limited the expansion of Tfh cells in the K/BxN mouse model of rheumatoid arthritis (RA). K/BxN mice treated with 2DG showed a reduced joint inflammation, which correlated with a decreased CD4+ T and B cell metabolism and a reduced activation of both adaptive and innate immune cells182. Therefore these results showed that autoreactive Tfh cells are highly dependent on glucose in multiple genetic backgrounds and with diverse autoantigens, effectively linking glycolysis, mTORC1 activation and Tfh expansion in autoantibody-mediated diseases. The requirement for mTORC1 activation to expand the number of autoreactive Tfh cells has been established in the Def6tr/trSwap70−/− DKO mice, in which a dysregulated T cell signaling leads to a lupus-like phenotype183. Interesting, at least in this model, the critical role of mTORC1 in Tfh cells is to promote the translation of Bcl6, the master regulator of Tfh cell gene expression184. This finding was unexpected since it has been shown that Bcl6 inhibits cellular metabolism including glycolysis, at least in vitro185. The same result was obtained in this study for the inhibitory receptor PD-1, which is highly expressed by Tfh cells185. Consistent with these findings, inhibiting glycolysis did not impair the induction of Tfh cells in response to protein immunization or to viral infection161. This suggests that autoreactive Tfh cells have a uniquely requirement for glucose, which may represent an Achilles’ heel for their selective elimination.

Treg cells do not require glucose for either their differentiation or suppressive function186. However, the migration of Treg cells to the site of inflammation, a process often overlooked but essential to Treg function in vivo, depends on glycolysis187. This process is mediated through a secondary role of the glycolytic enzyme glucokinase, whose expression is induced by PI3K-mTORC2 activation in Treg cells. Glucokinase binds actin and remodels the cytoskeleton allowing for cell migration. This result illustrates that specific metabolic pathways, such as here the PI3K/mTORC2 pathway, may sustain different functions of specific immune cells. In a systemic disease such as lupus in which multiple organs may be affected and which Treg cells are defective in both mouse models and patients188, treatment with glycolysis inhibitor may therefore have indirect detrimental consequences by increasing tissue inflammation.

Glucose is the main energy source during B cell activation and recent studies have uncovered context and stage specific glucose utilization during of B cell activation and differentiation189. B cells show increased glycolysis after activation by a range of stimuli, which result in HIF1α and c-Myc directly binding the promoter of genes encoding for glycolytic enzymes and glucose transporters in B cells190192. Directly relevant to lupus, high levels of BAFF, an essential B cell growth factor that is overexpressed in lupus patients, increased glycolysis in B cells190. B cell differentiation after antigen stimulation is energetically demanding and GC B cells increase their glucose consumption and mitochondria mass in comparison to naïve B cells158,193,194. The β isoform of protein kinase C (PKCβ), which is highly abundant in B cells and mediates proliferative signaling downstream of the B-cell receptor (BCR), is required for BCR-induced glycolysis195. Loss of PKCβ in B cells reduced the activation of the energy-regulating kinase complex mTORC1, resulting in a defective metabolism and mt remodeling, which impaired GC formation and the generation of plasma cells196. Plasma cells continuously utilize glucose once they are fully mature197. However, plasma cell survival and antibody production are maintained in the absence of Glut1, suggesting that other transporters are involved in glucose uptake190. Interestingly, glycolysis enhanced by hexokinase-2 overexpression increases plasma cell differentiation over self-renewal and long-term survival, while the inhibition of glycolysis impaired both plasma cell differentiation and survival198. Long-lived plasma cells (LLPCs) consume high amounts glucose, which is predominantly for antibody glycosylation. However, LLPCs divert some of this glucose towards pyruvate generation and respiration, a process that is required for their survival199. These results show that glucose is required for both lifespan and antibody secretion, although through different metabolic pathways. Surprisingly, glucose metabolism in B cells has not been directly examined in the context of lupus. Although a normalization of GC B cell phenotypes has been reported in lupus mice treated with 2DG161, it is unclear how much of this effect is B-cell intrinsic, or secondary to the effect of 2DG on Tfh cells.

Overall, a number of studies converge in identifying glycolysis as a pathway that may offer direct or indirect targets to limit the expansion of effector immune cells in autoimmune diseases leading to effective and maybe selective disease reduction. These promising results should however be validated in detailed disease-specific analyses of pre-clinical models to identify the specific pathways in the glucose flux and their cellular targets, as well as to what extent they are specific to autoreactive cells.

8. Oxidative phosphorylation and Kreb’s cycle in lupus

The ETC is a series of four multimeric protein complexes in the mt inner membrane in which electrons are transferred from donor to acceptors to ultimately produce ATP and reactive oxygen species (ROS) in a process known OXPHOS200. The electron donors NADH and FADH2 are produced by the Kreb’s cycle, which represents a chain of enzymatic reactions that oxidize fatty acids (FA), glutamine and acetyl-CoA derived from pyruvate. The Kreb’s cycle is also a source of metabolite intermediates for the synthesis of amino acids and FA. Truncated or “broken” Kreb’s cycles have been described in inflammatory conditions, leading to the accumulation of some of its metabolite intermediates, including ROS, instead of their recycling.

Type 1 IFN promotes OXPHOS and FA oxidation (FAO) in pDCs, which is critical for their activation and amplification of IFN production201. Since type 1 IFN production by pDCs is central to lupus pathogenesis, it is possible that OXPHOS inhibition specifically in pDCs would have beneficial effects. Although protocols to achieve cell-specific metabolic inhibition have not yet been identified, it is possible that the overactivation of a metabolic pathway in a specific cell type makes it a preferential target for inhibition, a hypothesis that needs to be formally tested. It is unknown whether other myeloid cells and B cells have altered OXPHOS or FAO metabolism in autoimmune diseases202. However, secondary signals delivered either by TLRs or T cell help are necessary to rescue BCR-activated B cells from mt dysfunction189. In addition, LLPCs, as stated above, require mt import of pyruvate199. Whether these metabolic determinants are altered in activated B cells and LLPCs in the context of lupus remains to be determined.

Multiple parameters of mt dysfunction has been reported in the CD4+ T cells from SLE patients and lupus-prone mice, including increased mt mass, membrane hyperpolarization, production of ROS intermediates and ATP depletion121. The ultimate cause of mt dysfunction resulting in a hyperoxidative state of CD4+ T cells in lupus is unknown, but may have a genetic basis. This is a largely unexplored hypothesis. However, in its support, the murine lupus gene corresponding to Sle1c2 is Esrrg, which encodes for the nuclear receptor ERRγ that transactivates the expression of many genes involved in mt metabolism, including OXPHOS, in many cell types203. The NZM2410 lupus-prone mice carry a hypomorph allele of Esrrg, which is associated with CD4+ T cells hyperactivation and defective Treg cells203.

Oxidative stress characterizes the lupus immune system and promotes its activation at multiple levels204. Oxidized mtDNA, which is released by necrotic cells as well as by the extracellular traps (NETs) produced by activated neutrophils, has emerged as a major stimulant of the production of type I IFN, and its circulating levels correlate with disease activity in SLE patients205,206. Moreover, CD4+ T cells of SLE patients are depleted in glutathione, a major natural antioxidant. Accordingly, the by-products of oxidative stress have been tested as therapeutic targets in lupus. A treatment with N-acetyl-cysteine (NAC) that replenishes glutathione levels normalized the elevated oxygen consumption by lupus CD4+ T cells in vitro207. Treatment of SLE patients with NAC decreased disease activity indexes in a double-blind placebo-controlled pilot study, with a reduction of mTOR activity in T cells208. Treatments with MitoTempo, another inhibitor of mt ROS production, delayed disease progression in MRL/lpr mice206, and it should be further explored in other preclinical models of lupus. Additional pre-clinical studies and clinical trials with anti-oxidants will be necessary to identify cellular targets and define biomarkers that may predict clinical responsiveness209.

A recent study revealed that kidney infiltrating T cells (KITs) in several models of lupus nephritis present a metabolically exhausted phenotype reminiscent to the exhaustion of tumor infiltrating lymphocytes210. OXPHOS and well as the spare respiratory capacity, which is defined as an energy reserve to be used upon challenge, were lost in both CD4+ and CD8+ KITs. It is unclear how metabolically exhausted T cells can participate to the disease process in lupus nephritis in mouse models211,212 and patients213,214. Another unexpected finding related to OXPHOS in lupus was that C1q, a member of the complement pathway that initiates efferocytosis, modulates the mt metabolism of CD8+ T cells in cGVHD, an induced model of SLE215. Mice deficient in C1q suffered a reduced cGVHD induction because C1q-deficient effector memory CD8+ T cells presented diminished mt functions. This novel role of C1q in the regulation of mt metabolism in CD8+ T cells needs to be further explored in the context of systemic autoimmunity in which little is known.

Although the mechanisms of mt dysfunction in lupus remain globally unexplored, its therapeutic potentials have been considered through the targeting of OXPHOS, glutaminolysis, FAO, and the Kreb’s cycle. Succinate, which classically bridges the Krebs cycle and OXPHOS, is also a mediator of inflammation216. Succinate builds up in Treg cells in which mt complex III has been deleted and contributes to their dysfunction217. In addition, a novel CXCR5CXCR3+PD1hiCD4+ T cell population provides help to B cells in the blood and the tubular interstitium of SLE patients in the form of succinate, along with IL-10, rather than IL-21218. Succinate accumulates in these helper T cells as a result of stimulation by oxidized mtDNA, high levels of ROS production and reverse electron transport, and it stabilizes HIF1α and increases glycolysis. Excess succinate secretion requires PD1 signaling in the CXCR5-negative helper T cells and binds the SUCNR1 succinate receptor cell expressed on B cells. SUCNR1 signaling was already known to induce pro-inflammatory phenotypes in innate immune cells, including RA219 and MetS220. This novel study218 now suggests that succinate sensing plays an essential role in the activation of autoreactive B cells in lupus. Itaconate is another metabolite that presents non-metabolic signaling functions. In addition to enhancing the oxidation of succinate to fumarate, and therefore reducing intracellular succinate accumulation, itaconate triggers multiple pathways to alleviate IL-17-driven inflammation221. Permeable itaconate, 4-octyl itaconate (OI) restricted the type 1 IFN response and the production of inflammatory cytokines by macrophages222,223. OI was tested in PBMCs from lupus patients in which it decreased the production of pro-inflammatory cytokines, such as TNFα and IL-6, through the activation of NRF2 signaling224. These important recent developments in immunometabolism research should lead the implementation of studies to test whether Kreb’s cycle metabolite intermediates that have secondary immune signaling functions could be targeted in lupus.

Metformin is a drug wide used to treat type 2 diabetes by preventing neoglucogenesis. Metformin transiently inhibits mt ETC complex I, decreasing OXPHOS and ATP production which indirectly leads to AMPK activation, with therefore the potential to interfere with key immunological processes. Metformin is anti-inflammatory by promoting the induction of Treg cell differentiation and blocking STAT3 activation, which has been attributed to either AMPK activation and mTORC1 inhibition, or increased FAO225. In vitro, metformin inhibits IFNγ production and promoted IL-2 production by CD4+ T cells from SLE patients and lupus-prone mice121. A treatment combining metformin and 2DG reversed disease biomarkers in several lupus mouse models, which correlated with a decreased respiration in CD4+ T cells121. Finally, metformin or other ETC inhibitors reduced the response of CD4+ T cells from both HCs and SLE patients to type 1 IFN by inhibiting pSTAT1 Y701 phosphorylation226. This result suggests that metformin treatment may benefit SLE patients with a high type 1 IFN activity. As mentioned above, metformin added to standard-of-care treatment reduced the risk of disease flares and corticosteroid exposure in SLE patients with mild/moderate disease activity120. The mechanisms of action of metformin were not investigated in this study. Nonetheless, these promising results have led to a prospective, multicenter, double blinded, placebo controlled, clinical trial ( NCT02741960) of metformin as an add-on therapy in SLE patients with mild disease activity.

Finally, ATP, the final product of ETC activity, has an immunoregulatory role beyond its primary metabolic function. Dying cells release ATP, which is transported back into cells by the P2X7 ATP receptor227. Deletion of P2X7 exacerbated autoimmune pathology in the pristane-induced model of lupus228. This phenotype was associated with an expanded number of Tfh cells and GC B cells. Conversely, treatment with a P2X7 pharmacological agonist (BzATP) reversed this process. Interestingly, the generation on foreign-antigen specific Tfh cells was not impaired by P2X7 deficiency. In accordance with these findings in mice, P2X7 expression and function are reduced in the PBMCs of lupus patients as compared to HCs229. These results suggest that P2X7 may provide another metabolic therapeutic target to reduce the number of pathogenic cell types or redirect their function in lupus. These results also support the hypothesis that autoreactive Tfh cells have unique metabolic requirements, which are fueled by glucose (see above), and dampened by extra-cellular ATP.

9. Fatty acid oxidation and lupus

Long chain free FAs are imported into immune cells by members of the SLC27 family and CD36, to be then incorporated into FA acetyl-CoA by long-chain FA-CoA ligase. FA acetyl-CoA are carried into mt by the carnitine shuttle, which is the rate-limiting step in FA catabolism. The two-carbon (acetate) units are then removed through β oxidation and combined with co-enzyme A (co-A) to form acetyl-coA, which enters the Kreb’s cycle. Etomoxir, an inhibitor of CPT1α, the enzyme that controls the carnitine shuttle, has been used to investigate FAO in immune cells. The role of FAO in any immune cells type has not yet been characterized in the context of autoimmune diseases. A number of findings potentially relevant to autoimmunity and lupus have however been obtained with non-autoimmune cells. Since Treg and memory T cells use FAO as their main energy source, etomoxir treatment of non-autoimmune mice impaired Treg cell differentiation and function while activated inflammatory T cell subsets remained untouched186. In inflammatory conditions represented by the GVHD model, different results were obtained with the most striking effect of etomoxir occurred in effector T cells, by suppressed alloreactive T cells while other T cell populations remained unperturbed230. Although the relative role of FAO relative to glycolysis or oxidation of glucose or glutamine has not been defined in these alloreactive T cells, these results suggest that inhibiting FAO could limit the activation of CD4+ T cells in the context of autoimmune pathogenesis, including lupus. Caution should be used however in the interpretation of the results obtained from mice or cells commonly treated with supra-physiological doses of etomoxir, which deplete the pool of free coA, a central metabolite for the Kreb’s cycle, FA synthesis, and histone acetylation231,232. Ideally, results obtained with etomoxir should be validated by a genetic approach directly targeting CPT1α.

Indirectly linked to FAO, the metabolism of the sugar D-mannose modulates immune responses by inducing Treg cell differentiation and decreasing the production of inflammatory cytokines233. Glycolysis is inhibited by D-mannose, forcing immune cells to switch to FAO, which generates mtROS, which, in turn, promotes TGF-β production with its well documented immunoregulatory effects233. These results illustrates how metabolic interventions should be interpreted carefully with a full assessment of complex interconnected metabolic networks, which may have unexpected consequences on immune activation. This study also suggests that D-mannose may be considered in pre-clinical models of autoimmunity in which Treg cell expansion may have a therapeutic benefit.

While the role of FAO has been mostly characterized in T cells, at least in vitro, FAO inhibition also prevented TLR9-induced activation of both conventional DCs and pDCs234. Although the production of type I IFN was not directly measured, the expression of IFN-induced CXCL10 was decreased by FAO inhibition in these cells. This study also found that pDCs were more sensitive than cDCs to FAO inhibition, which is consistent with pDCs heavily relying on OXPHOS and FAO for activation201. These results suggest that FAO could be a promising therapeutic target in lupus by reducing the pathogenicity of T cells and pDCs. However, its involvement in Treg cell maintenance and function would probably raise concerns. Another limitation to targeting FAO is the lack of small molecule inhibitors that can be used in pre-clinical models.

10. Branch chain amino acid and glutamine metabolism

In addition to being protein building blocks, specific amino acids are at the basis of many anabolic pathways, including the synthesis of lipids, nucleotides, glutathione, glucosamine, and polyamines. Furthermore, glutamine (Gln) is directly used to produce energy through the anaplerosis of the Kreb’s cycle235. Amino acid synthesis also plays a major part in immune activation, as simply illustrated by the dramatic increase in amino acid levels, including Gln in CD4+ T cells activated in vitro through their receptor and CD28236. Moreover, branched chain amino acids (BCAA) such as leucine and Gln function as metabolic sensors of a cell energy status by directly activating mTORC1237. Tfh cells in lupus-prone mice display a specific solute carrier expression signature161, which includes several amino acid transporters such as Slc7a5, Slc7a10, ASCT2 and LAT1/CD98. While the functional significance of this differential expression is currently unknown, it is increasingly recognized that a better understanding of the functional link between solute transporters in general, these amino acid transporters in particular, and the metabolic programing of immune cells, will unlock novel regulatory circuits of immune activation238. The lack of reagents, such as antibodies, inhibitors, and cell-specific deletions for many Slc members are still a major hurdle toward this goal.

Leucine metabolism contributes to autoimmune pathology as a checkpoint of mTORC1 signaling, which then controls glycolysis with the consequences described above on effector T cells and myeloid cells239241. Leucine is transported into cells by SLC7A5, and subjected to a series of enzymatic reactions to produce to acetyl-CoA, which will enter the Kreb’s cycle. The first rate-limiting reaction is a reversible transamination of leucine to α-ketoisocaproate, mediated by branched-chain aminotransferase (BCAT). There are two isoforms of BCAT based on their cellular location, mt BCAT2 and cytosolic BCAT1. BCAT1 is the most abundantly expressed BCAT isoform in human macrophages242. The inhibition of BCAT1 presented beneficial effects in pre-clinical models of RA and crescentic glomerulonephritis by reducing the producing of inflammatory cytokines by macrophages and reducing their infiltration in target organs242. This suggests that the consumption of leucine by the TCA cycle is pro-inflammatory in macrophages in inflammatory conditions.

Gln imported by ASCT2 / SLC1A5 is converted first to glutamate by glutaminase (GLS), then to α-ketoglutaric acid (αKG) by transaminases or GLUD1. αKG enters the Kreb’s cycle or is converted by isocitrate dehydrogenases (IDH1 and IDH2) into (D)-2-hydroxyglutarate (2HG), which inhibits DNA and histone demethylases. This latter pathway places Gln metabolism as an essential gate to epigenetic regulation243. Glutaminolysis is a major determinant of the balance between Th17 and Treg cell differentiation, which may define its major role in autoimmune diseases. Gln is a major energy source for the Th17 cells7. Gln deprivation in the culture media promoted Treg differentiation from either naïve CD4+ T cells7 or Th1-polarized cells244. Another study reported that Gln depletion achieved either by transporter deficiency or deprivation in the culture media inhibited both Th1 and Th17 differentiation245. GLS inhibition however preferentially inhibited Th17 polarization7,246. Gls deficiency in T cells increased T-bet expression as well as the differentiation and effector functions of Th1 and CD8+ cytotoxic cells, but these phenotypes were unstable, indicating that glutaminolysis was required for their optimal functions. Th17 cell differentiation was severely impaired in Gls deficient T cells247. The inhibition of 2HG production through glutamate oxaloacetate transaminase 1 (GOT1) also blocked Th17 polarization in favor of Treg cells but did not affect Th1 cells248. Thus, there is mounting evidence that glutaminolysis is essential for Th17 cells. The inconsistent findings between studies regarding the Gln requirements of Th1 cells may be due to differences in experimental conditions or indicate that glutaminolysis in Th1 may be compensated by cytokines such as IL-2 or by yet undefined anaplerotic sources in certain conditions. In the context of lupus, Gls1 inhibition with BPTES ameliorated lupus manifestations in MRL/lpr mice in a Th17-dependent manner15. It is unknown whether this treatment would be as effective in other models in which Th17 cells play a lesser role. The requirements for glutaminolysis have been recently examined for the GC reaction in a mouse model of lupus. Contrary to glycolysis, glutaminolysis inhibition with the Gls1 inhibitor 6-diazo-5-oxo-l-norleucine (DON) greatly reduced immunization-induced as well as autoimmune GC responses in both lupus-prone and non-autoimmune mice161. GC B cells were nearly eliminated and Tfh cells presented a reduced expression of a number of genes associated with their function, including Bcl6. These results indicate that glutaminolysis is an absolute metabolic checkpoint for the development of all GCs. Lower levels of Gln were found in PBMCs from lupus patients than HCs, with an inverse correlation between Gln levels and disease activity249. This study concluded that Gln deficiency may contribute to mt dysfunction in SLE patients. This result may also be interpreted as SLE patients having a higher consumption of Gln, and it would be of great interest to determine whether it correlates with the frequency of specific T cell subsets in these patients.

Finally, as mentioned above, glutaminolysis regulates DNA demethylation. This epigenetic regulation of gene expression is produced through the conversion of 5-methylcytosine to 5-hydroxymethylcytosine by enzymes belonging to the TET family. This oxidation reaction occurs through the engagement of glutaminolysis metabolites, such as α-KG and 2-HG. High Gln utilization in differentiating Th17 cells results in an accumulation of 2-HG, which inhibits DNA demethylases, maintaining the Foxp3 locus hypermethylated and silenced248. This process explains the bi-directional effect of glutaminolysis on Th17 and Treg cells. Despite an increased interest to understand the complex epigenetic regulation of autoimmune activation, including in lupus250, Gln-driven epigenetic modifications have not been directly examined in this context. These studies should be a part of mechanistic investigations following up the promising results showing a protective effect of Gln inhibition.

Multiple studies have targeted Gln metabolism in an attempt to starve off tumors. Successful strategies include the inhibition of Gln transporters SLC1A5/ASCT2 and SLC7A11/LAT1 with V-9302 or GPNA, and with BCH or xCT system inhibitors respectively; blocking the first reaction of glutaminolysis with GLS1 inhibitors DON, CB-839, 968, or BPTES; or blocking glutamate dehydrogenase (GDH) with EGCG, and aminotransferase with AOA251. These drugs represent a promising tool box to use in pre-clinical models of autoimmunity, including in lupus. Glutaminolysis is a complex pathway that affects multiple cellular processes with far-reaching consequences that are still largely unexplored in autoimmunity. As illustrated by studies reviewed above, many steps from the Gln transporters to the many enzymatic reactions represent potential targets to treat lupus. Pre-clinical and mechanistic studies need to be carefully conducted to identify targets as well as long-term consequences on the entire immune system to avoid immunosuppression. A recent study has however generated a lot of enthusiasm by showing that tumor infiltrating T cells and the tumor cells have opposite Gln requirements allowing for a treatment with DON to simultaneously activate cytotoxic T cells and starve the tumor252.

11. Lipid synthesis

FA synthesis is initiated by the conversion of acetyl-CoA into malonyl-CoA by the rate-limiting enzyme ACC1253. Subsequent steps are performed by fatty acid synthase (FASN), stearoyl-CoA desaturase (SCD), and the FA-coenzyme A ligase family to generate diacetyl- and triacetyl-glycerols and long-chained FA254. The majority of de novo synthesized FA are then directed to the plasma membrane incorporated into phospholipids or form lipid rafts to regulate the clustering of membrane-anchored receptors that is trigger signaling in immune cells255.

Triglycerides, phosphoglycerides, or sphingolipids are forms for FA that directly regulate T cell responses not only as key components of cell membranes, but also as signaling molecules, and energy sources. The inhibition of FA synthesis in T cells by ACC1 deletion resulted in defective blasting and massive activation-induced cell death that prevented the induction of an antigen-specific CD8+ T cell response. Interestingly, this defect was rescued by exogenous FA, indicating that it was the abundance, but not the source of FAs that controls the survival of antigen-activated CD8+ T cells256. These results suggest that ACC1 may be a therapeutic target to modulate CD8+ T cell activity. As mentioned above, the role of CD8+ T cells in lupus is not well understood, but they may contribute to organ damage by direct cell killing, and therefore ACC1 inhibition may be beneficial. Acc1 expression and the levels of activated phosphorylated ACC1 increased during Th17 cell differentiation. The functional consequence of this observation was demonstrated by the pharmacological inhibition or T cell-specific deletion of ACC1, which inhibited the polarization of human and murine Th17 cells in favor of Treg cell induction not only in vitro, but also in an EAE mouse model257. Similar to ACC1, FASN inhibition also reduced Th17 cell polarization, but it also uniquely boosted IFN-γ production by Th1 and Th1-like Th17 cells258. Again, it is not clear whether these differences are due to experimental differences or if they reflects the different roles of the metabolites produced by the two enzymatic reactions. Furthermore, it is still unknown how FA synthesis regulates CD8+ T cell survival or inflammatory CD4+ T cell polarization. Its profound effect on the Th17 / Treg balance is however reminiscent of Gln metabolism, and, as such, it is predicted to regulate lupus pathogenesis. However, the paucity of FA synthesis inhibitors suitable for in vivo treatments makes it a difficult hypothesis to test.

Glycosphingolipids (GSLs) are a major component of lipid rafts. A combination of increased synthesis and reduced recycling results in high levels of GSL in the CD4+ T cells from SLE patients, which were directed by an increased LXRβ expression259. Pharmacological inhibition of GSL synthesis with N-butyldeoxynojirimycin (Miglustat), a clinically approved drug used to treat Gaucher disease, normalized the function in the lupus T cells, including their ability to induce anti-DNA antibody production by syngeneic B cells. An independent study confirmed the critical role of GSL in lupus T cells. The capacity of BTLA, an inhibitory receptor similar in function to CTLA-4 and PD-1, to restrain T cell activation is defective in SLE patients260. This is at least in part due to a poor BTLA recruitment to the immunological synapse, as shown by a treatment with a glucosylceramide synthase inhibitor, which normalized lipid metabolism as well as BTLA inhibitory function. Mechanistically, decreased glucosylceramide availability dissociated lipid rafts clustered around TCR molecules, allowing BTLA recruitment and inhibition of TCR signaling260. These two studies convincingly linked GSL homeostasis to the strength of T cell activation in lupus, and warrant follow studies in animal models to assess in vivo therapeutic effects.

Finally, FA also form cytoplasmic lipid droplets261, which have been found necessary for T cells to invade the joints of RA patients, in which the inhibition of FA synthesis prevented this process262. It has not been examined whether T cells that invade the inflamed organs of lupus patients, including the joints and the kidneys, are also dependent on de novo FA synthesis.

12. Conclusions and perspectives

Research in immunometabolism is a growing field that started with in vitro studies that proposed simple models in which effector immune cells switched upon activation from mt metabolism to aerobic glycolysis that was supported by a large increase in glucose uptake263. It has now been established that Gln and FA utilization also represent critical checkpoints of metabolic reprogramming regulating immune cell functions. Metabolic inhibitors and gene targeting of metabolic enzymes have been widely used to extend the reach of these findings mostly in mouse models, but also with some human cells. The vast majority of results were obtained with healthy mice and cells from healthy subjects. As reported in this review, a number of immunophenotypes directly relevant to autoimmunity can be manipulated through their metabolism, which should be directly tested in animal models, as well as in cells from patients with autoimmune diseases. As to be expected, multiple levels of complexity have been added to the initial models of metabolic reprograming controlling immune functions, which should be considered if immunometabolism is to be targeted for therapeutic purposes.

Many aspects of metabolic reprogramming may be specific to an immune cell type, with the most striking example being pDC activation by type I IFN relying on FAO201, a unique process not found in other types of immune cells switching to glycolysis upon activation. A corollary of the cell specificity of metabolic pathways is that targeting one pathway to eliminate or expand the number of a specific cell type could be detrimental to another cell type also involved in disease. This is an issue technically difficult to address since beneficial effects of successful interventions with metabolic inhibitors have been correlated with changes in given cell populations, such as Th17 or Treg cells, but the direct cellular targets are unknown. Metabolic alterations may also be disease-specific, with stark differences even between two rheumatic diseases such as RA and SLE that present significant etiological overlaps. Indeed, antioxidants are beneficial for CD4+ T cells in SLE while oxidative agents eliminate the pathogenicity of CD4+ T cells in RA160,264. Therefore, it is unfortunate but critically important to not assume that the findings obtained for one autoimmune disease can be automatically translated to another autoimmune disease.

Metabolism is known for its intricate complexity at the biochemical level. Recent studies have added levels of complexity showing “moonlighting” functions of some metabolic enzymes, such as GAPDDH regulating IFNγ production265, and metabolites, such as succinate or itaconate with previously unsuspected immune signaling functions266. Finally, mitochondria have been solidly emerged as the command center of immune functions through an increasing number of mechanisms, not only as the major source of ATP production, but only through the generation of these metabolite intermediates including ROS production, and by providing a platform for RIG-I and NRLP-3 inflammasome signaling. The highly immunogenic mtDNA leading to type I IFN production267.suggests that mt hyperoxidation is a key pathogenic feature of lupus. Therefore, the restoration of mt health is predicted to have a major therapeutic effect in lupus patients, but an effective treatment to achieve this goal has not yet been found.

Recent developments suggest that research in immunometabolism will lead to long-lasting impactful discoveries in the field of autoimmunity, including lupus. First, lupus is the first autoimmune disease in which metabolic inhibitors, mTOR inhibitors, NAC and metformin, have shown therapeutic benefits in patients and as well as others in several pre-clinical models (Table 1). Additional pathways and inhibitors should be tested based on their efficacy on inflammatory immune cells elicited in healthy mice. Second, a disease-specific cell-specific understanding of the metabolic alterations at the molecular level may provide novel mechanistic insights on autoimmune activation, which may in turn provide much needed novel therapeutic targets. Third, immunometabolism studies should explore whether metabolic inhibitors have therapeutic value added to standard-of-care treatments and to biologics that have not met endpoints in lupus clinical trials. Immunometabolism is already the target of methotrexate, a drug widely used in rheumatology that inhibits 1-Carbon metabolism268. The pre-clinical studies proposed this review conceived and interpreted in the framework of results obtained in normal mice, coupled with a deep characterization of the metabolic signatures of effector cells in mice and patients with autoimmune diseases should be prioritized to reap the therapeutic promises of metabolic targeting.

Acknowledgments:

This work is supported by grants to LM from the NIH (R01 AI045050 and R01 AI128901) and from the Alliance for Lupus Research (550197)

Footnotes

Conflict of interest: None

References

  • 1.Frauwirth KA, Riley JL, Harris MH, et al. The CD28 signaling pathway regulates glucose metabolism. Immunity. 2002;16:769–777. [DOI] [PubMed] [Google Scholar]
  • 2.Doughty CA, Bleiman BF, Wagner DJ, et al. Antigen receptor–mediated changes in glucose metabolism in B lymphocytes: role of phosphatidylinositol 3-kinase signaling in the glycolytic control of growth. Blood. 2006;107:4458–4465. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Buck MD, Sowell RT, Kaech SM, Pearce EL. Metabolic instruction of immunity. Cell. 2017;169:570–586. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.O’Sullivan D, Pearce EL. Targeting T cell metabolism for therapy. Trends Immunol. 2015;36:71–80. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Rhoads JP, Major AS, Rathmell JC. Fine tuning of immunometabolism for the treatment of rheumatic diseases. Nat Rev Rheumatol. 2017;13:313–320. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Vander Heiden MG, Cantley LC, Thompson CB. Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science. 2009;324:1029–1033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Sugiura A, Rathmell JC. Metabolic barriers to T cell function in tumors. J Immunol. 2018;200:400–407. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Vander Heiden MG, DeBerardinis RJ. Understanding the intersections between metabolism and cancer biology. Cell. 2017;168:657–669. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Perl A. Metabolic control of immune system activation in rheumatic diseases. Arthritis Rheumatol. 2017;69:2259–2270. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Morel L. Immunometabolism in systemic lupus erythematosus. Nat Rev Rheumatol. 2017;13:280–290. [DOI] [PubMed] [Google Scholar]
  • 11.Liu E, Perl A. Pathogenesis and treatment of autoimmune rheumatic diseases. Curr Opin Rheumatol. 2019;31:307–315. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Teng X, Cornaby C, Li W, Morel L. Metabolic regulation of pathogenic autoimmunity: therapeutic targeting. Curr Opin Immunol. 2019;61:10–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Zhang D, Jin W, Wu R, et al. High glucose intake exacerbates autoimmunity through reactive-oxygen-species-mediated TGF-beta cytokine activation. Immunity. 2019;51:671–681 e675. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Shi LZ, Wang R, Huang G, et al. HIF1α–dependent glycolytic pathway orchestrates a metabolic checkpoint for the differentiation of TH17 and Treg cells. J Exp Med. 2011;208:1367–1376. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Kono M, Yoshida N, Maeda K, Suarez-Fueyo A, Kyttaris VC, Tsokos GC. Glutaminase 1 inhibition reduces glycolysis and ameliorates lupus-like disease in MRL/lpr mice and experimental autoimmune encephalomyelitis. Arthritis Rheumatol. 2019;71:1869–1878. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Sun C, Qin W, Zhang YH, et al. Prevalence and risk of metabolic syndrome in patients with systemic lupus erythematosus: A meta-analysis. Int J Rheum Dis. 2017;20:917–928. [DOI] [PubMed] [Google Scholar]
  • 17.Hallajzadeh J, Khoramdad M, Izadi N, et al. The association between metabolic syndrome and its components with systemic lupus erythematosus: a comprehensive systematic review and meta-analysis of observational studies. Lupus. 2018:961203317751047. [DOI] [PubMed] [Google Scholar]
  • 18.Parker B, Ahmad Y, Shelmerdine J, et al. An analysis of the metabolic syndrome phenotype in systemic lupus erythematosus. Lupus. 2011;20:1459–1465. [DOI] [PubMed] [Google Scholar]
  • 19.Gami AS, Witt BJ, Howard DE, et al. Metabolic syndrome and risk of incident cardiovascular events and death: a systematic review and meta-analysis of longitudinal studies. J Am Coll Cardiol. 2007;49:403–414. [DOI] [PubMed] [Google Scholar]
  • 20.Mok CC, Tse SM, Chan KL, Ho LY. Effect of the metabolic syndrome on organ damage and mortality in patients with systemic lupus erythematosus: a longitudinal analysis. Clin Exp Rheumatol. 2018;36:389–395. [PubMed] [Google Scholar]
  • 21.Demir S, Erten G, Artım-Esen B, et al. Increased serum leptin levels are associated with metabolic syndrome and carotid intima media thickness in premenopausal systemic lupus erythematosus patients without clinical atherosclerotic vascular events. Lupus. 2018;27:1509–1516. [DOI] [PubMed] [Google Scholar]
  • 22.Tedeschi SK, Barbhaiya M, Malspeis S, et al. Obesity and the risk of systemic lupus erythematosus among women in the Nurses’ Health Studies. Semin Arthritis Rheum. 2017;47:376–383. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Cozier YC, Barbhaiya M, Castro-Webb N, et al. A prospective study of obesity and risk of systemic lupus erythematosus (SLE) among Black women. Semin Arthritis Rheum. 2019;48:1030–1034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Yuan J, Li LI, Wang Z, Song W, Zhang Z. Dyslipidemia in patients with systemic lupus erythematosus: Association with disease activity and B-type natriuretic peptide levels. Biomed Rep. 2016;4:68–72. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Ryu H, Lim H, Choi G, et al. Atherogenic dyslipidemia promotes autoimmune follicular helper T cell responses via IL-27. Nat Immunol. 2018;19:583–593. [DOI] [PubMed] [Google Scholar]
  • 26.Choi JY, Ho JH, Pasoto SG, et al. Circulating follicular helper-like T cells in systemic lupus erythematosus: association with disease activity. Arthritis Rheumatol. 2015;67:988–999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Ryu H, Lim H, Choi G, et al. Atherogenic dyslipidemia promotes autoimmune follicular helper T cell responses via IL-27. Nat Immunol. 2018;19:583–593. [DOI] [PubMed] [Google Scholar]
  • 28.Lee MH, Gallo PM, Hooper KM, et al. The cytokine network type I IFN-IL-27-IL-10 is augmented in murine and human lupus. J Leukoc Biol. 2019;106:967–975. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Wilhelm AJ, Zabalawi M, Grayson JM, et al. Apolipoprotein A-I and its role in lymphocyte cholesterol homeostasis and autoimmunity. Arterioscler Thromb Vasc Biol. 2009;29:843–849. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Abella V, Scotece M, Conde J, et al. Leptin in the interplay of inflammation, metabolism and immune system disorders. Nat Rev Rheumatol. 2017;13:100–109. [DOI] [PubMed] [Google Scholar]
  • 31.Collins N, Han SJ, Enamorado M, et al. The bone marrow protects and optimizes immunological memory during dietary restriction. Cell. 2019;178:1088–1101.e1015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Nagai M, Noguchi R, Takahashi D, et al. Fasting-Refeeding Impacts Immune Cell Dynamics and Mucosal Immune Responses. Cell. 2019;178:1072–1087.e1014. [DOI] [PubMed] [Google Scholar]
  • 33.Jordan S, Tung N, Casanova-Acebes M, et al. Dietary intake regulates the circulating inflammatory monocyte pool. Cell. 2019;178:1102–1114.e1117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Choi IY, Piccio L, Childress P, et al. A diet mimicking fasting promotes regeneration and reduces autoimmunity and multiple sclerosis symptoms. Cell Rep. 2016;15:2136–2146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Cignarella F, Cantoni C, Ghezzi L, et al. Intermittent fasting confers protection in cns autoimmunity by altering the gut microbiota. Cell Metab. 2018;27:1222–1235.e1226. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Rangan P, Choi I, Wei M, et al. Fasting-mimicking diet modulates microbiota and promotes intestinal regeneration to reduce inflammatory bowel disease pathology. Cell Rep. 2019;26:2704–2719.e2706. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Sun D, Krishnan A, Su J, Lawrence R, Zaman K, Fernandes G. Regulation of immune function by calorie restriction and cyclophosphamide treatment in lupus-prone NZB/NZW F1 mice. Cell Immunol. 2004;228:54–65. [DOI] [PubMed] [Google Scholar]
  • 38.Fernandes G, Yunis EJ, Good RA. Influence of diet on survival of mice. Proc Natl Acad Sci USA. 1976;73:1279–1283. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Fernandes G, Friend P, Yunis EJ, Good RA. Influence of dietary restriction on immunologic function and renal disease in (NZB × NZW) F1 mice. Proc Natl Acad Sci USA. 1978;75:1500–1504. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Mittal A, Muthukumar A, Jolly CA, Zaman K, Fernandes G. Reduced food consumption increases water intake and modulates renal aquaporin-1 and −2 expression in autoimmune prone mice. Life Sci. 2000;66:1471–1479. [DOI] [PubMed] [Google Scholar]
  • 41.He J, Yang B. Aquaporins in renal diseases. Int J Mol Sci. 2019;20 pii: E366. doi: 10.3390/ijms20020366 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Troyer DA, Chandrasekar B, Barnes JL, Fernandes G. Calorie restriction decreases platelet-derived growth factor (PDGF)-A and thrombin receptor mRNA expression in autoimmune murine lupus nephritis. Clin Exp Immunol. 1997;108:58–62. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Muthukumar AR, Jolly CA, Zaman K, Fernandes G. Calorie restriction decreases proinflammatory cytokines and polymeric Ig receptor expression in the submandibular glands of autoimmune prone (NZB × NZW)F1 mice. J Clin Immunol. 2000;20:354–361. [DOI] [PubMed] [Google Scholar]
  • 44.Davies RJ, Lomer MC, Yeo SI, Avloniti K, Sangle SR, D’Cruz DP. Weight loss and improvements in fatigue in systemic lupus erythematosus: a controlled trial of a low glycaemic index diet versus a calorie restricted diet in patients treated with corticosteroids. Lupus. 2012;21:649–655. [DOI] [PubMed] [Google Scholar]
  • 45.Madeira I, Bordallo MA, Rodrigues NC, et al. Leptin as a predictor of metabolic syndrome in prepubertal children. Arch Endocrinol Metab. 2017;61:7–13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.La Cava A, Matarese G. The weight of leptin in immunity. Nat Rev Immunol. 2004;4:371–379. [DOI] [PubMed] [Google Scholar]
  • 47.Garcia-Gonzalez A, Gonzalez-Lopez L, Valera-Gonzalez IC, et al. Serum leptin levels in women with systemic lupus erythematosus. Rheumatol Int. 2002;22:138–141. [DOI] [PubMed] [Google Scholar]
  • 48.Wang X, Qiao Y, Yang L, et al. Leptin levels in patients with systemic lupus erythematosus inversely correlate with regulatory T cell frequency. Lupus. 2017;26:1401–1406. [DOI] [PubMed] [Google Scholar]
  • 49.Lee YH, Song GG. Association between circulating leptin levels and systemic lupus erythematosus: an updated meta-analysis. Lupus. 2018;27:428–435. [DOI] [PubMed] [Google Scholar]
  • 50.Diaz-Rizo V, Bonilla-Lara D, Gonzalez-Lopez L, et al. Serum levels of adiponectin and leptin as biomarkers of proteinuria in lupus nephritis. PLoS One. 2017;12:e0184056. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Kim HA, Choi GS, Jeon JY, Yoon JM, Sung JM, Suh CH. Leptin and ghrelin in Korean systemic lupus erythematosus. Lupus. 2010;19:170–174. [DOI] [PubMed] [Google Scholar]
  • 52.Liu Y, Yu Y, Matarese G, La Cava A. Cutting edge: fasting-induced hypoleptinemia expands functional regulatory T cells in systemic lupus erythematosus. J Immunol. 2012;188:2070–2073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Chougule D, Nadkar M, Venkataraman K, et al. Adipokine interactions promote the pathogenesis of systemic lupus erythematosus. Cytokine. 2018;111:20–27. [DOI] [PubMed] [Google Scholar]
  • 54.De Sanctis JB, Zabaleta M, Bianco NE, Garmendia JV, Rivas L. Serum adipokine levels in patients with systemic lupus erythematosus. Autoimmunity. 2009;42:272–274. [DOI] [PubMed] [Google Scholar]
  • 55.Toussirot E, Gaugler B, Bouhaddi M, Nguyen NU, Saas P, Dumoulin G. Elevated adiponectin serum levels in women with systemic autoimmune diseases. Mediators Inflamm. 2010;2010:938408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Li HM, Zhang TP, Leng RX, et al. Association of leptin and leptin receptor gene polymorphisms with systemic lupus erythematosus in a Chinese population. J Cell Mol Med. 2017;21:1732–1741. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Vadacca M, Zardi EM, Margiotta D, et al. Leptin, adiponectin and vascular stiffness parameters in women with systemic lupus erythematosus. Intern Emerg Med. 2013;8:705–712. [DOI] [PubMed] [Google Scholar]
  • 58.McMahon M, Skaggs BJ, Sahakian L, et al. High plasma leptin levels confer increased risk of atherosclerosis in women with systemic lupus erythematosus, and are associated with inflammatory oxidised lipids. Ann Rheum Dis. 2011;70:1619–1624. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Reis BS, Lee K, Fanok MH, et al. Leptin receptor signaling in T cells is required for Th17 differentiation. J Immunol. 2015;194:5253–5260. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Lord GM, Matarese G, Howard JK, Baker RJ, Bloom SR, Lechler RI. Leptin modulates the T-cell immune response and reverses starvation-induced immunosuppression. Nature. 1998;394:897–901. [DOI] [PubMed] [Google Scholar]
  • 61.Farooqi IS, Matarese G, Lord GM, et al. Beneficial effects of leptin on obesity, T cell hyporesponsiveness, and neuroendocrine/metabolic dysfunction of human congenital leptin deficiency. J Clin Invest. 2002;110:1093–1103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Chen J, Zeng W, Pan W, et al. Symptoms of systemic lupus erythematosus are diagnosed in leptin transgenic pigs. PLoS Biol. 2018;16:e2005354. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Fujita Y, Fujii T, Mimori T, et al. Deficient leptin signaling ameliorates systemic lupus erythematosus lesions in MRL/Mp-Fas lpr mice. J Immunol. 2014;192:979–984. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Lourenço EV, Liu A, Matarese G, La Cava A. Leptin promotes systemic lupus erythematosus by increasing autoantibody production and inhibiting immune regulation. Proc Natl Acad Sci USA. 2016;113:10637–10642. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Yu Y, Liu Y, Shi FD, Zou H, Matarese G, La Cava A. Cutting edge: Leptin-induced RORγt expression in CD4+ T cells promotes Th17 responses in systemic lupus erythematosus. J Immunol. 2013;190:3054–3058. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Gerriets VA, Danzaki K, Kishton RJ, et al. Leptin directly promotes T-cell glycolytic metabolism to drive effector T-cell differentiation in a mouse model of autoimmunity. Eur J Immunol. 2016;46:1970–1983. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Gupta S, Agrawal S, Gollapudi S. Increased activation and cytokine secretion in B cells stimulated with leptin in aged humans. Immun Ageing. 2013;10:3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Agrawal S, Gollapudi S, Su H, Gupta S. Leptin activates human B cells to secrete TNF-α, IL-6, and IL-10 via JAK2/STAT3 and p38MAPK/ERK1/2 signaling pathway. J Clin Immunol. 2011;31:472–478. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Amarilyo G, Iikuni N, Shi FD, Liu A, Matarese G, La Cava A. Leptin promotes lupus T-cell autoimmunity. Clin Immunol. 2013;149:530–533. [DOI] [PubMed] [Google Scholar]
  • 70.Amarilyo G, Iikuni N, Liu A, Matarese G, La Cava A. Leptin enhances availability of apoptotic cell-derived self-antigen in systemic lupus erythematosus. PLoS One. 2014;9:e112826. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Fang WL, Zhou B, Wang YY, Chen Y, Zhang L. Analysis of adiponectin gene polymorphisms in Chinese population with systemic lupus erythematosus. J Biomed Biotechnol. 2010;2010:401537. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Loghman M, Haghighi A, Broumand B, et al. Association between urinary adiponectin level and renal involvement in systemic lupus erythematous. Int J Rheum Dis. 2016;19:678–684. [DOI] [PubMed] [Google Scholar]
  • 73.Mahieu MA, Ahn GE, Chmiel JS, et al. Serum adipokine levels and associations with patient-reported fatigue in systemic lupus erythematosus. Rheumatol Int. 2018;38:1053–1061. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Dini AA, Wang P, Ye DQ. Serum Adiponectin Levels in Patients With Systemic Lupus Erythematosus: A Meta-analysis. J Clin Rheumatol. 2017;23:361–367. [DOI] [PubMed] [Google Scholar]
  • 75.Zhang TP, Zhao YL, Li XM, Wu CH, Pan HF, Ye DQ. Altered mRNA expression levels of vaspin and adiponectin in peripheral blood mononuclear cells of systemic lupus erythematosus patients. Clin Exp Rheumatol. 2019;37:458–464. [PubMed] [Google Scholar]
  • 76.Parker J, Menn-Josephy H, Laskow B, Takemura Y, Aprahamian T. Modulation of lupus phenotype by adiponectin deficiency in autoimmune mouse models. J Clin Immunol. 2011;31:167–173. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Ito A, Hong C, Oka K, et al. Cholesterol accumulation in CD11c(+) immune cells is a causal and targetable factor in autoimmune disease. Immunity. 2016;45:1311–1326. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Hiltbold EM, Poloso NJ, Roche PA. MHC class II-peptide complexes and APC lipid rafts accumulate at the immunological synapse. J Immunol. 2003;170:1329–1338. [DOI] [PubMed] [Google Scholar]
  • 79.Wang SH, Yuan SG, Peng DQ, Zhao SP. HDL and ApoA-I inhibit antigen presentation-mediated T cell activation by disrupting lipid rafts in antigen presenting cells. Atherosclerosis. 2012;225:105–114. [DOI] [PubMed] [Google Scholar]
  • 80.Bonacina F, Coe D, Wang G, et al. Myeloid apolipoprotein E controls dendritic cell antigen presentation and T cell activation. Nat Commun. 2018;9:3083. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Eberle D, Hegarty B, Bossard P, Ferre P, Foufelle F. SREBP transcription factors: master regulators of lipid homeostasis. Biochimie. 2004;86:839–848. [DOI] [PubMed] [Google Scholar]
  • 82.York AG, Williams KJ, Argus JP, et al. Limiting cholesterol biosynthetic flux spontaneously engages type I IFN signaling. Cell. 2015;163:1716–1729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Han S, Zhuang H, Shumyak S, et al. Liver X receptor agonist therapy prevents diffuse alveolar hemorrhage in murine lupus by repolarizing macrophages. Front Immunol. 2018;9:135. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.A-Gonzalez N, Bensinger SJ, Hong C, et al. Apoptotic cells promote their own clearance and immune tolerance through activation of the nuclear receptor LXR. Immunity. 2009;31:245–258. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Westerterp M, Gourion-Arsiquaud S, Murphy AJ, et al. Regulation of hematopoietic stem and progenitor cell mobilization by cholesterol efflux pathways. Cell Stem Cell. 2012;11:195–206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Westerterp M, Gautier EL, Ganda A, et al. Cholesterol accumulation in dendritic cells links the inflammasome to acquired immunity. Cell Metab. 2017;25:1294–1304 e1296. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Yvan-Charvet L, Wang N, Tall AR. Role of HDL, ABCA1, and ABCG1 transporters in cholesterol efflux and immune responses. Arterioscler Thromb Vasc Biol. 2010;30:139–143. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Armstrong AJ, Gebre AK, Parks JS, Hedrick CC. ATP-binding cassette transporter G1 negatively regulates thymocyte and peripheral lymphocyte proliferation. J Immunol. 2010;184:173–183. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Bengtsson AA, Trygg J, Wuttge DM, et al. Metabolic profiling of systemic lupus erythematosus and comparison with primary Sjogren’s syndrome and systemic sclerosis. PloS one. 2016;11:e0159384. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Wu T, Xie C, Han J, et al. Metabolic disturbances associated with systemic lupus erythematosus. PLoS One. 2012;7:e37210. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Ouyang X, Dai Y, Wen JL, Wang LX. ¹H NMR-based metabolomic study of metabolic profiling for systemic lupus erythematosus. Lupus. 2011;20:1411–1420. [DOI] [PubMed] [Google Scholar]
  • 92.Guleria A, Pratap A, Dubey D, et al. NMR based serum metabolomics reveals a distinctive signature in patients with Lupus Nephritis. Sci Rep. 2016;6:35309. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Scavuzzi BM, Simão ANC, Iriyoda TMV, et al. Increased lipid and protein oxidation and lowered anti-oxidant defenses in systemic lupus erythematosus are associated with severity of illness, autoimmunity, increased adhesion molecules, and Th1 and Th17 immune shift. Immunol Res. 2018;66:158–171. [DOI] [PubMed] [Google Scholar]
  • 94.Akesson K, Pettersson S, Stahl S, et al. Kynurenine pathway is altered in patients with SLE and associated with severe fatigue. Lupus Sci Med. 2018;5:e000254. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Lood C, Tydén H, Gullstrand B, et al. Type I interferon-mediated skewing of the serotonin synthesis is associated with severe disease in systemic lupus erythematosus. PLoS One. 2015;10:e0125109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Widner B, Sepp N, Kowald E, Kind S, Schmuth M, Fuchs D. Degradation of tryptophan in patients with systemic lupus erythematosus. Adv Exp Med Biol. 1999;467:571–577. [DOI] [PubMed] [Google Scholar]
  • 97.Widner B, Sepp N, Kowald E, et al. Enhanced tryptophan degradation in systemic lupus erythematosus. Immunobiol. 2000;201:621–630. [DOI] [PubMed] [Google Scholar]
  • 98.Mandel EH, Appleton HD. Tryptophan metabolism. Results of studies in discoid lupus erythematosus. Arch Dermatol. 1966;94:358–360. [DOI] [PubMed] [Google Scholar]
  • 99.Pertovaara M, Hasan T, Raitala A, et al. Indoleamine 2,3-dioxygenase activity is increased in patients with systemic lupus erythematosus and predicts disease activation in the sunny season. Clin Exp Immunol. 2007;150:274–278. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Perl A, Hanczko R, Lai ZW, et al. Comprehensive metabolome analyses reveal. Metabolomics. 2015;11:1157–1174. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Yan B, Huang J, Zhang C, et al. Serum metabolomic profiling in patients with systemic lupus erythematosus by GC/MS. Mod Rheumatol. 2016;26:914–922. [DOI] [PubMed] [Google Scholar]
  • 102.Shin TH, Kim HA, Jung JY, et al. Analysis of the free fatty acid metabolome in the plasma of patients with systemic lupus erythematosus and fever. Metabolomics. 2017;14:14. [DOI] [PubMed] [Google Scholar]
  • 103.Urowitz MB, Gladman D, Ibañez D, et al. Clinical manifestations and coronary artery disease risk factors at diagnosis of systemic lupus erythematosus: data from an international inception cohort. Lupus. 2007;16:731–735. [DOI] [PubMed] [Google Scholar]
  • 104.Müller K, Kriegbaum NJ, Baslund B, Sørensen OH, Thymann M, Bentzen K. Vitamin D3 metabolism in patients with rheumatic diseases: low serum levels of 25-hydroxyvitamin D3 in patients with systemic lupus erythematosus. Clin Rheumatol. 1995;14:397–400. [DOI] [PubMed] [Google Scholar]
  • 105.Mok CC, Birmingham DJ, Leung HW, Hebert LA, Song H, Rovin BH. Vitamin D levels in Chinese patients with systemic lupus erythematosus: relationship with disease activity, vascular risk factors and atherosclerosis. Rheumatology (Oxford). 2012;51:644–652. [DOI] [PubMed] [Google Scholar]
  • 106.Bogaczewicz J, Sysa-Jedrzejowska A, Arkuszewska C, et al. Vitamin D status in systemic lupus erythematosus patients and its association with selected clinical and laboratory parameters. Lupus. 2012;21:477–484. [DOI] [PubMed] [Google Scholar]
  • 107.Mok CC, Birmingham DJ, Ho LY, Hebert LA, Song H, Rovin BH. Vitamin D deficiency as marker for disease activity and damage in systemic lupus erythematosus: a comparison with anti-dsDNA and anti-C1q. Lupus. 2012;21:36–42. [DOI] [PubMed] [Google Scholar]
  • 108.Zhang Q, Yin X, Wang H, et al. Fecal metabolomics and potential biomarkers for systemic lupus erythematosus. Front Immunol. 2019;10:976. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Hevia A, Milani C, López P, et al. Intestinal dysbiosis associated with systemic lupus erythematosus. MBio. 2014;5:e01548–01514. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Luo XM, Edwards MR, Mu Q, et al. Gut microbiota in human systemic lupus erythematosus and a mouse model of lupus. Appl Environ Microbiol. 2018;84. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.He Z, Shao T, Li H, Xie Z, Wen C. Alterations of the gut microbiome in Chinese patients with systemic lupus erythematosus. Gut Pathog. 2016;8:64. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Rodríguez-Carrio J, López P, Sánchez B, et al. Intestinal dysbiosis is associated with altered short-chain fatty acids and serum-free fatty acids in systemic lupus erythematosus. Front Immunol. 2017;8:23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Azzouz D, Omarbekova A, Heguy A, et al. Lupus nephritis is linked to disease-activity associated expansions and immunity to a gut commensal. Ann Rheum Dis. 2019;78:947–956. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Li Y, Wang H, Li X, et al. Disordered intestinal microbes are associated with the activity of Systemic Lupus Erythematosus. Clin Sci (Lond). 2019;133:821–838. [DOI] [PubMed] [Google Scholar]
  • 115.Ma Y, Xu X, Li M, Cai J, Wei Q, Niu H. Gut microbiota promote the inflammatory response in the pathogenesis of systemic lupus erythematosus. Mol Med. 2019;25:35. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Yan B, Huang J, Dong F, et al. Urinary metabolomic study of systemic lupus erythematosus based on gas chromatography/mass spectrometry. Biomed Chromatogr. 2016;30:1877–1881. [DOI] [PubMed] [Google Scholar]
  • 117.Powell JD, Pollizzi KN, Heikamp EB, Horton MR. Regulation of immune responses by mTOR. Ann Rev Immunol. 2012;30:39–68. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Inoki K, Kim J, Guan KL. AMPK and mTOR in cellular energy homeostasis and drug targets. Annu Rev Pharmacol Toxicol. 2012;52:381–400. [DOI] [PubMed] [Google Scholar]
  • 119.Zhou G, Myers R, Li Y, et al. Role of AMP-activated protein kinase in mechanism of metformin action. J Clin Invest. 2001;108:1167–1174. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Wang H, Li T, Chen S, Gu Y, Ye S. NETs mitochondrial DNA and its autoantibody in Systemic Lupus Erythematosus and a proof-of-concept trial of metformin. Arthritis Rheumatol. 2015;67:3190–3200. [DOI] [PubMed] [Google Scholar]
  • 121.Yin Y, Choi SC, Xu Z, et al. Normalization of CD4+ T cell metabolism reverses lupus. Sci Transl Med. 2015;7:274ra218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Yin Y, Choi S-C, Xu Z, et al. Glucose oxidation is critical for CD4+ T cell activation in a mouse model of systemic lupus erythematosus. J Immunol. 2016;196:80–90. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Lee SY, Moon SJ, Kim EK, et al. Metformin Suppresses systemic autoimmunity in roquin(san/san) mice through inhibiting b cell differentiation into plasma cells via regulation of AMPK/mTOR/STAT3. J Immunol. 2017;198:2661–2670. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Fernandez D, Perl A. mTOR signaling: a central pathway to pathogenesis in systemic lupus erythematosus? Discov Med. 2010;9:173–178. [PMC free article] [PubMed] [Google Scholar]
  • 125.Lui SL, Tsang R, Chan KW, et al. Rapamycin attenuates the severity of established nephritis in lupus-prone NZB/W F1 mice. Nephrol Dialysis Transplant. 2008;23:2768–2776. [DOI] [PubMed] [Google Scholar]
  • 126.Lai ZW, Kelly R, Winans T, et al. Sirolimus in patients with clinically active systemic lupus erythematosus resistant to, or intolerant of, conventional medications: a single-arm, open-label, phase 1/2 trial. Lancet. 2018;391:1186–1196. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Yap DYH, Tang C, Chan GCW, et al. Longterm data on sirolimus treatment in patients with lupus nephritis. J Rheumatol. 2018;45:1663–1670. [DOI] [PubMed] [Google Scholar]
  • 128.Delgoffe GM, Kole TP, Zheng Y, et al. The mTOR kinase differentially regulates effector and regulatory T cell lineage commitment. Immunity. 2009;30:832–844. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Zeng H, Cohen S, Guy C, et al. mTORC1 and mTORC2 kinase signaling and glucose metabolism drive follicular helper T cell differentiation. Immunity. 2016;45:540–554. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Suarez-Fueyo A, Bradley SJ, Tsokos GC. T cells in Systemic Lupus Erythematosus. Curr Opin Immunol. 2016;43:32–38. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Pratama A, Vinuesa CG. Control of TFH cell numbers: why and how? Immunol Cell Biol. 2014;92:40–48. [DOI] [PubMed] [Google Scholar]
  • 132.Yang K, Blanco DB, Neale G, et al. Homeostatic control of metabolic and functional fitness of Treg cells by LKB1 signalling. Nature. 2017;548:602–606. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.He N, Fan W, Henriquez B, et al. Metabolic control of regulatory T cell (Treg) survival and function by Lkb1. Proc Natl Acad Sci USA. 2017;114:12542–12547. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Sun IH, Oh MH, Zhao L, et al. mTOR complex 1 signaling regulates the generation and function of central and effector Foxp3(+) regulatory T cells. J Immunol. 2018;201:481–492. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Chapman NM, Zeng H, Nguyen TM, et al. mTOR coordinates transcriptional programs and mitochondrial metabolism of activated Treg subsets to protect tissue homeostasis. Nat Commun. 2018;9:2095. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Katsiari CG, Kyttaris VC, Juang YT, Tsokos GC. Protein phosphatase 2A is a negative regulator of IL-2 production in patients with systemic lupus erythematosus. J Clin Invest. 2005;115:3193–3204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Apostolidis SA, Rodriguez-Rodriguez N, Suarez-Fueyo A, et al. Phosphatase PP2A is requisite for the function of regulatory T cells. Nat Immunol. 2016;17:556–564. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Xu B, Wang S, Zhou M, et al. The ratio of circulating follicular T helper cell to follicular T regulatory cell is correlated with disease activity in systemic lupus erythematosus. Clin Immunol. 2017;183:46–53. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Xu L, Huang Q, Wang H, et al. The kinase mTORC1 promotes the generation and suppressive function of follicular regulatory T cells. Immunity. 2017;47:538–551 e535. [DOI] [PubMed] [Google Scholar]
  • 140.Torigoe M, Iwata S, Nakayamada S, et al. Metabolic reprogramming commits differentiation of human CD27(+)IgD(+) B cells to plasmablasts or CD27(−)IgD(−) cells. J Immunol. 2017;199:425–434. [DOI] [PubMed] [Google Scholar]
  • 141.Iwata TN, Ramirez-Komo JA, Park H, Iritani BM. Control of B lymphocyte development and functions by the mTOR signaling pathways. Cytokine Growth Factor Rev. 2017;35:47–62. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Iwata TN, Ramirez JA, Tsang M, et al. Conditional disruption of raptor reveals an essential role for mTORC1 in B cell development, survival, and metabolism. J Immunol. 2016;197:2250–2260. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Farmer JR, Allard-Chamard H, Sun N, et al. Induction of metabolic quiescence defines the transitional to follicular B cell switch. Sci Signal. 2019;12:eaaw5573. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Keating R, Hertz T, Wehenkel M, et al. The kinase mTOR modulates the antibody response to provide cross-protective immunity to lethal infection with influenza virus. Nat Immunol. 2013;14:1266–1276. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Zhang S, Pruitt M, Tran D, et al. B cell-specific deficiencies in mTOR limit humoral immune responses. J Immunol. 2013;191:1692–1703. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Jones DD, Gaudette BT, Wilmore JR, et al. mTOR has distinct functions in generating versus sustaining humoral immunity. J Clin Invest. 2016;126:4250–4261. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Goldfinger M, Shmuel M, Benhamron S, Tirosh B. Protein synthesis in plasma cells is regulated by crosstalk between endoplasmic reticulum stress and mTOR signaling. Eur J Immunol. 2011;41:491–502. [DOI] [PubMed] [Google Scholar]
  • 148.Pengo N, Scolari M, Oliva L, et al. Plasma cells require autophagy for sustainable immunoglobulin production. Nat Immunol. 2013;14:298–305. [DOI] [PubMed] [Google Scholar]
  • 149.Hiepe F, Radbruch A. Plasma cells as an innovative target in autoimmune disease with renal manifestations. Nat Rev Nephrol. 2016;12:232–240. [DOI] [PubMed] [Google Scholar]
  • 150.Thoreen CC, Chantranupong L, Keys HR, Wang T, Gray NS, Sabatini DM. A unifying model for mTORC1-mediated regulation of mRNA translation. Nature. 2012;485:109–113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Lee K, Heffington L, Jellusova J, et al. Requirement for Rictor in homeostasis and function of mature B lymphoid cells. Blood. 2013;122:2369–2379. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Sukhbaatar N, Hengstschlager M, Weichhart T. mTOR-mediated regulation of dendritic cell differentiation and function. Trends Immunol. 2016;37:778–789. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Corcoran SE, O’Neill LA. HIF1alpha and metabolic reprogramming in inflammation. The J Clin Invest. 2016;126:3699–3707. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Shi LZ, Wang R, Huang G, et al. HIF1alpha-dependent glycolytic pathway orchestrates a metabolic checkpoint for the differentiation of TH17 and Treg cells. J Exp Med. 2011;208:1367–1376. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Yao Y, Wang L, Zhou J, Zhang X. HIF-1alpha inhibitor echinomycin reduces acute graft-versus-host disease and preserves graft-versus-leukemia effect. J Transl Med. 2017;15:28. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Davidson A. What is damaging the kidney in lupus nephritis? Nat Rev Rheumatol. 2016;12:143–153. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Cho SH, Raybuck AL, Stengel K, et al. Germinal centre hypoxia and regulation of antibody qualities by a hypoxia response system. Nature. 2016;537:234–238. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Jellusova J, Cato MH, Apgar JR, et al. Gsk3 is a metabolic checkpoint regulator in B cells. Nat Immunol. 2017;18:303–312. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Cho SH, Raybuck AL, Blagih J, et al. Hypoxia-inducible factors in CD4(+) T cells promote metabolism, switch cytokine secretion, and T cell help in humoral immunity. Proc Nat Acad Sci USA. 2019;116:8975–8984. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Perl A. Oxidative stress in the pathology and treatment of systemic lupus erythematosus. Nat Rev Rheumatol. 2013;9:674–686. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Choi SC, Titov AA, Abboud G, et al. Inhibition of glucose metabolism selectively targets autoreactive follicular helper T cells. Nat Commun. 2018;9:4369. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Macintyre AN, Gerriets VA, Nichols AG, et al. The glucose transporter Glut1 is selectively essential for CD4 T cell activation and effector function. Cell Metab. 2014;20:61–72. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Jacobs SR, Herman CE, Maciver NJ, et al. Glucose uptake is limiting in T cell activation and requires CD28-mediated Akt-dependent and independent pathways. J Immunol. 2008;180:4476–4486. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Koga T, Sato T, Furukawa K, et al. Promotion of calcium/calmodulin-dependent protein kinase 4 by GLUT1-dependent glycolysis in systemic lupus erythematosus. Arthritis Rheumatol. 2019;71:766–772. [DOI] [PubMed] [Google Scholar]
  • 165.Ferretti AP, Bhargava R, Dahan S, Tsokos MG, Tsokos GC. Calcium/calmodulin kinase IV controls the function of both T cells and kidney resident cells. Front Immunol. 2018;9:2113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Koga T, Hedrich CM, Mizui M, et al. CaMK4-dependent activation of AKT/mTOR and CREM-alpha underlies autoimmunity-associated Th17 imbalance. J Clin Invest. 2014;124:2234–2245. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Sena LA, Li S, Jairaman A, et al. Mitochondria are required for antigen-specific T cell activation through reactive oxygen species signaling. Immunity. 2013;38:225–236. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Zhang D, Li J, Wang F, Hu J, Wang S, Sun Y. 2-Deoxy-D-glucose targeting of glucose metabolism in cancer cells as a potential therapy. Cancer letters. 2014;355:176–183. [DOI] [PubMed] [Google Scholar]
  • 169.Nguyen HD, Chatterjee S, Haarberg KM, et al. Metabolic reprogramming of alloantigen-activated T cells after hematopoietic cell transplantation. J Clin Invest. 2016;126:1337–1352. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Shirai T, Nazarewicz RR, Wallis BB, et al. The glycolytic enzyme PKM2 bridges metabolic and inflammatory dysfunction in coronary artery disease. J Exp Med. 2016;213:337–354. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Kono M, Maeda K, Stocton-Gavanescu I, et al. Pyruvate kinase M2 is requisite for Th1 and Th17 differentiation. JCI Insight. 2019;4 pii: 127395. doi: 10.1172/jci.insight.127395 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Angiari S, Runtsch MC, Sutton CE, et al. Pharmacological activation of pyruvate kinase M2 inhibits CD4(+) T cell pathogenicity and suppresses autoimmunity. Cell Metab. 2019. pii: S1550–4131(19)30606–0. doi: 10.1016/j.cmet.2019.10.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Kono M, Yoshida N, Maeda K, et al. Pyruvate dehydrogenase phosphatase catalytic subunit 2 limits Th17 differentiation. Proc Nat Acad Sci USA. 2018;115:9288–9293. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Kornberg MD, Bhargava P, Kim PM, et al. Dimethyl fumarate targets GAPDH and aerobic glycolysis to modulate immunity. Science. 2018;360:449–453. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Deeks ED. Dimethyl fumarate: A review in relapsing-remitting MS. Drugs. 2016;76:243–254. [DOI] [PubMed] [Google Scholar]
  • 176.Kuhn A, Landmann A, Patsinakidis N, et al. Fumaric acid ester treatment in cutaneous lupus erythematosus (CLE): a prospective, open-label, phase II pilot study. Lupus. 2016;25:1357–1364. [DOI] [PubMed] [Google Scholar]
  • 177.Saracino AM, Orteu CH. Severe recalcitrant cutaneous manifestations in systemic lupus erythematosus successfully treated with fumaric acid esters. Brit J Dermatol. 2017;176:472–480. [DOI] [PubMed] [Google Scholar]
  • 178.Di Dedda C, Vignali D, Piemonti L, Monti P. Pharmacological targeting of GLUT1 to control autoreactive t cell responses. Int J Mol Sci. 2019;20:4962. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Li W, Qu G, Choi SC, et al. Targeting T cell activation and lupus autoimmune phenotypes by inhibiting glucose transporters. Front Immunol. 2019;10:833. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180.MacIver NJ, Michalek RD, Rathmell JC. Metabolic regulation of T lymphocytes. Ann Rev Immunol. 2013;31:259–283. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Ma EH, Verway MJ, Johnson RM, et al. Metabolic profiling using stable isotope tracing reveals distinct patterns of glucose utilization by physiologically activated CD8(+) T cells. Immunity. 2019;51:856–870 e855. [DOI] [PubMed] [Google Scholar]
  • 182.Abboud G, Choi SC, Kanda N, Zeumer-Spataro L, Roopenian DC, Morel L. Inhibition of Glycolysis Reduces Disease Severity in an Autoimmune Model of Rheumatoid Arthritis. Front Immunol. 2018;9:1973. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Yi W, Gupta S, Ricker E, et al. The mTORC1–4E-BP-eIF4E axis controls de novo Bcl6 protein synthesis in T cells and systemic autoimmunity. Nat Commun. 2017;8:254. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Patsoukis N, Bardhan K, Chatterjee P, et al. PD-1 alters T-cell metabolic reprogramming by inhibiting glycolysis and promoting lipolysis and fatty acid oxidation. Nat Commun. 2015;6:6692. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Oestreich KJ, Read KA, Gilbertson SE, et al. Bcl-6 directly represses the gene program of the glycolysis pathway. Nat Immunol. 2014;15:957–964. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Michalek RD, Gerriets VA, Jacobs SR, et al. Cutting edge: distinct glycolytic and lipid oxidative metabolic programs are essential for effector and regulatory CD4+ T cell subsets. J Immunol. 2011;186:3299–3303. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Kishore M, Cheung KCP, Fu H, et al. Regulatory T cell migration is dependent on glucokinase-mediated glycolysis. Immunity. 2017;47:875–889 e810. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Ohl K, Tenbrock K. Regulatory T cells in systemic lupus erythematosus. Eur J Immunol. 2015;45:344–355. [DOI] [PubMed] [Google Scholar]
  • 189.Akkaya M, Pierce SK. From zero to sixty and back to zero again: the metabolic life of B cells. Curr Opin Immunol. 2019;57:1–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Caro-Maldonado A, Wang R, Nichols AG, et al. Metabolic reprogramming is required for antibody production that is suppressed in anergic but exaggerated in chronically BAFF-exposed B cells. J Immunol. 2014;192:3626–3636. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Doughty CA, Bleiman BF, Wagner DJ, et al. Antigen receptor-mediated changes in glucose metabolism in B lymphocytes: role of phosphatidylinositol 3-kinase signaling in the glycolytic control of growth. Blood. 2006;107:4458–4465. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Dufort FJ, Bleiman BF, Gumina MR, et al. Cutting edge: IL-4-mediated protection of primary B lymphocytes from apoptosis via Stat6-dependent regulation of glycolytic metabolism. J Immunol. 2007;179:4953–4957. [DOI] [PubMed] [Google Scholar]
  • 193.Ersching J, Efeyan A, Mesin L, et al. Germinal center selection and affinity maturation require dynamic regulation of mTORC1 kinase. Immunity. 2017;46:1045–1058 e1046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Jayachandran N, Mejia EM, Sheikholeslami K, et al. TAPP adaptors control B cell metabolism by modulating the phosphatidylinositol 3-kinase signaling pathway: a novel regulatory circuit preventing autoimmunity. J Immunol. 2018;201:406–416. [DOI] [PubMed] [Google Scholar]
  • 195.Blair D, Dufort FJ, Chiles TC. Protein kinase Cbeta is critical for the metabolic switch to glycolysis following B-cell antigen receptor engagement. Biochem J. 2012;448:165–169. [DOI] [PubMed] [Google Scholar]
  • 196.Tsui C, Martinez-Martin N, Gaya M, et al. Protein kinase C-beta dictates B cell fate by regulating mitochondrial remodeling, metabolic reprogramming, and heme biosynthesis. Immunity. 2018;48:1144–1159 e1145. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Lam WY, Bhattacharya D. Metabolic Links between plasma cell survival, secretion, and stress. Trends Immunol. 2018;39:19–27. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198.Adams WC, Chen YH, Kratchmarov R, et al. Anabolism-associated mitochondrial stasis driving lymphocyte differentiation over self-renewal. Cell Rep. 2016;17:3142–3152. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Lam WY, Becker AM, Kennerly KM, et al. Mitochondrial pyruvate import promotes long-term survival of antibody-secreting plasma cells. immunity. 2016;45:60–73. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Smeitink J, van den Heuvel L, DiMauro S. The genetics and pathology of oxidative phosphorylation. Nat Rev Genet. 2001;2:342–352. [DOI] [PubMed] [Google Scholar]
  • 201.Wu D, Sanin David E, Everts B, et al. Type 1 interferons induce changes in core metabolism that are critical for immune function. Immunity. 2016;44:1325–1336. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Waters LR, Ahsan FM, Wolf DM, Shirihai O, Teitell MA. Initial B cell activation induces metabolic reprogramming and mitochondrial remodeling. iScience. 2018;5:99–109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203.Perry DJ, Yin Y, Telarico T, et al. Murine lupus susceptibility locus Sle1c2 mediates CD4+ T cell activation and maps to estrogen-related receptor gamma. J Immunol. 2012;189:793–803. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 204.Morel L. Immunometabolism in systemic lupus erythematosus. Nat Rev Rheumatol. 2017;13:280–290. [DOI] [PubMed] [Google Scholar]
  • 205.Caielli S, Athale S, Domic B, et al. Oxidized mitochondrial nucleoids released by neutrophils drive type I interferon production in human lupus. J Exp Med. 2016;213:697–713. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Lood C, Blanco LP, Purmalek MM, et al. Neutrophil extracellular traps enriched in oxidized mitochondrial DNA are interferogenic and contribute to lupus-like disease. Nat Med. 2016;22:146–153. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207.Doherty E, Oaks Z, Perl A. Increased mitochondrial electron transport chain activity at complex I is regulated by N-acetylcysteine in lymphocytes of patients with systemic lupus erythematosus. Antioxid Redox Signal. 2014;21:56–65. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208.Lai ZW, Hanczko R, Bonilla E, et al. N-acetylcysteine reduces disease activity by blocking mammalian target of rapamycin in T cells from systemic lupus erythematosus patients: a randomized, double-blind, placebo-controlled trial. Arthritis Rheum. 2012;64:2937–2946. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209.Perl A, Hanczko R, Lai ZW, et al. Comprehensive metabolome analyses reveal N-acetylcysteine-responsive accumulation of kynurenine in systemic lupus erythematosus: implications for activation of the mechanistic target of rapamycin. Metabolomics. 2015;11:1157–1174. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210.Tilstra JS, Avery L, Menk AV, et al. Kidney-infiltrating T cells in murine lupus nephritis are metabolically and functionally exhausted. J Clin Invest. 2018;128:4884–4897. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211.Odegard JM, Marks BR, DiPlacido LD, et al. ICOS-dependent extrafollicular helper T cells elicit IgG production via IL-21 in systemic autoimmunity. J Exp Med. 2008;205:2873–2886. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Moulton VR, Tsokos GC. Abnormalities of T cell signaling in systemic lupus erythematosus. Arthritis Res Ther. 2011;13:207. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213.Massengill SF, Goodenow MM, Sleasman JW. SLE nephritis is associated with an oligoclonal expansion of intrarenal T cells. Am J Kidney Dis. 1998;31:418–426. [DOI] [PubMed] [Google Scholar]
  • 214.Winchester R, Wiesendanger M, Zhang HZ, et al. Immunologic characteristics of intrarenal T cells: trafficking of expanded CD8+ T cell beta-chain clonotypes in progressive lupus nephritis. Arthritis Rheum. 2012;64:1589–1600. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Ling GS, Crawford G, Buang N, et al. C1q restrains autoimmunity and viral infection by regulating CD8(+) T cell metabolism. Science. 2018;360:558–563. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.Mills EL, Kelly B, Logan A, et al. Succinate dehydrogenase supports metabolic repurposing of mitochondria to drive inflammatory macrophages. Cell. 2016;167:457–470 e413. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217.Weinberg SE, Singer BD, Steinert EM, et al. Mitochondrial complex III is essential for suppressive function of regulatory T cells. Nature. 2019;565:495–499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Caielli S, Veiga DT, Balasubramanian P, et al. A CD4(+) T cell population expanded in lupus blood provides B cell help through interleukin-10 and succinate. Nat Med. 2019;25:75–81. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 219.Saraiva AL, Veras FP, Peres RS, et al. Succinate receptor deficiency attenuates arthritis by reducing dendritic cell traffic and expansion of Th17 cells in the lymph nodes. FASEB J. 2018;32:fj201800285. [DOI] [PubMed] [Google Scholar]
  • 220.Keiran N, Ceperuelo-Mallafre V, Calvo E, et al. SUCNR1 controls an anti-inflammatory program in macrophages to regulate the metabolic response to obesity. Nat Immunol. 2019;20:581–592. [DOI] [PubMed] [Google Scholar]
  • 221.Bambouskova M, Gorvel L, Lampropoulou V, et al. Electrophilic properties of itaconate and derivatives regulate the IkappaBzeta-ATF3 inflammatory axis. Nature. 2018;556:501–504. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Mills EL, Ryan DG, Prag HA, et al. Itaconate is an anti-inflammatory metabolite that activates Nrf2 via alkylation of KEAP1. Nature. 2018;556:113–117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223.Nonnenmacher Y, Hiller K. Biochemistry of proinflammatory macrophage activation. Cell Mol Life Sci. 2018;75:2093–2109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 224.Tang C, Wang X, Xie Y, et al. 4-Octyl itaconate activates Nrf2 signaling to inhibit pro-inflammatory cytokine production in peripheral blood mononuclear cells of systemic lupus erythematosus patients. Cell Physiol Biochem. 2018;51:979–990. [DOI] [PubMed] [Google Scholar]
  • 225.Lee SY, Lee SH, Yang EJ, et al. Metformin ameliorates inflammatory bowel disease by suppression of the STAT3 signaling pathway and regulation of the between Th17/Treg balance. PloS one. 2015;10:e0135858. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 226.Titov AA, Baker HV, Brusko TM, Sobel ES, Morel L. Metformin inhibits the type 1 IFN response in human CD4(+) T cells. J Immunol. 2019:ji1801651. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227.Di Virgilio F, Dal Ben D, Sarti AC, Giuliani AL, Falzoni S. The P2X7 Receptor in Infection and Inflammation. Immunity. 2017;47:15–31. [DOI] [PubMed] [Google Scholar]
  • 228.Faliti CE, Gualtierotti R, Rottoli E, et al. P2X7 receptor restrains pathogenic Tfh cell generation in systemic lupus erythematosus. J Exp Med. 2019;216:317–336. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 229.Furini F, Giuliani AL, Parlati ME, Govoni M, Di Virgilio F, Bortoluzzi A. P2X7 receptor expression in patients with serositis related to systemic lupus erythematosus. Front Pharmacol. 2019;10:435. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 230.Byersdorfer CA, Tkachev V, Opipari AW, et al. Effector T cells require fatty acid metabolism during murine graft-versus-host disease. Blood. 2013;122:3230–3237. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 231.Divakaruni AS, Hsieh WY, Minarrieta L, et al. Etomoxir inhibits macrophage polarization by disrupting coA homeostasis. Cell Metab. 2018;28:490–503 e497. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 232.Raud B, Roy DG, Divakaruni AS, et al. Etomoxir actions on regulatory and memory t cells are independent of Cpt1a-mediated fatty acid oxidation. Cell Metab. 2018;28:504–515 e507. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 233.Zhang D, Chia C, Jiao X, et al. D-mannose induces regulatory T cells and suppresses immunopathology. Nat Med. 2017;23:1036–1045. [DOI] [PubMed] [Google Scholar]
  • 234.Qiu CC, Atencio AE, Gallucci S. Inhibition of fatty acid metabolism by etomoxir or TOFA suppresses murine dendritic cell activation without affecting viability. Immunopharmacol Immunotoxicol. 2019;41:361–369. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 235.Tsun ZY, Possemato R. Amino acid management in cancer. Semin Cell Dev Biol. 2015;43:22–32. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 236.Wang R, Dillon C-P, Shi L-Z, et al. The transcription factor Myc controls metabolic reprogramming upon T lymphocyte activation. Immunity. 2011;35:871–882. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 237.Shimobayashi M, Hall MN. Multiple amino acid sensing inputs to mTORC1. Cell Res. 2016;26:7–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 238.Hsu C-L, Dzhagalov IL. Metabolite transporters—The Gatekeepers for T cell metabolism. Immunometabolism. 2019;1:e190012. [Google Scholar]
  • 239.Yoon BR, Oh YJ, Kang SW, Lee EB, Lee WW. Role of SLC7A5 in metabolic reprogramming of human monocyte/macrophage immune responses. Front Immunol. 2018;9:53. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 240.Ren W, Liu G, Yin J, et al. Amino-acid transporters in T-cell activation and differentiation. Cell Death Dis. 2017;8:e2655. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 241.Sinclair LV, Rolf J, Emslie E, Shi YB, Taylor PM, Cantrell DA. Control of amino-acid transport by antigen receptors coordinates the metabolic reprogramming essential for T cell differentiation. Nat Immunol. 2013;14:500–508. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 242.Papathanassiu AE, Ko JH, Imprialou M, et al. BCAT1 controls metabolic reprogramming in activated human macrophages and is associated with inflammatory diseases. Nat Comm. 2017;8:16040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 243.Wang Z, Long H, Chang C, Zhao M, Lu Q. Crosstalk between metabolism and epigenetic modifications in autoimmune diseases: a comprehensive overview. Cell Mol Life Sci. 2018;75:3353–3369. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 244.Klysz D, Tai X, Robert PA, et al. Glutamine-dependent alpha-ketoglutarate production regulates the balance between T helper 1 cell and regulatory T cell generation. Sci Signal. 2015;8:ra97. [DOI] [PubMed] [Google Scholar]
  • 245.Nakaya M, Xiao Y, Zhou X, et al. Inflammatory T cell responses rely on amino acid transporter ASCT2 facilitation of glutamine uptake and mTORC1 kinase activation. Immunity. 2014;40:692–705. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 246.Kono M, Yoshida N, Maeda K, Tsokos GC. Transcriptional factor ICER promotes glutaminolysis and the generation of Th17 cells. Proc Nat Acad Sci USA. 2018;115:2478–2483. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 247.Johnson MO, Wolf MM, Madden MZ, et al. Distinct regulation of Th17 and Th1 cell differentiation by glutaminase-dependent metabolism. Cell. 2018;175:1780–1795 e1719. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 248.Xu T, Stewart KM, Wang X, et al. Metabolic control of TH17 and induced Treg cell balance by an epigenetic mechanism. Nature. 2017;548:228–233. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 249.Lee HT, Lin CS, Pan SC, et al. Alterations of oxygen consumption and extracellular acidification rates by glutamine in PBMCs of SLE patients. Mitochondrion. 2019;44:65–74. [DOI] [PubMed] [Google Scholar]
  • 250.Wu H, Chen Y, Zhu H, Zhao M, Lu Q. The pathogenic role of dysregulated epigenetic modifications in autoimmune diseases. Front Immunol. 2019;10:2305. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 251.Li L, Meng Y, Li Z, et al. Discovery and development of small molecule modulators targeting glutamine metabolism. Eur J Med Chem. 2019;163:215–242. [DOI] [PubMed] [Google Scholar]
  • 252.Leone RD, Zhao L, Englert JM, et al. Glutamine blockade induces divergent metabolic programs to overcome tumor immune evasion. Science. 2019;366:1013–1021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 253.Wellen KE, Thompson CB. A two-way street: reciprocal regulation of metabolism and signalling. Nat Rev Mol Cell Biol. 2012;13:270–276. [DOI] [PubMed] [Google Scholar]
  • 254.Menendez JA, Lupu R. Fatty acid synthase and the lipogenic phenotype in cancer pathogenesis. Nat Rev Cancer. 2007;7:763–777. [DOI] [PubMed] [Google Scholar]
  • 255.Marat AL, Haucke V. Phosphatidylinositol 3-phosphates-at the interface between cell signalling and membrane traffic. EMBO J. 2016;35:561–579. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 256.Lee J, Walsh MC, Hoehn KL, James DE, Wherry EJ, Choi Y. Regulator of fatty acid metabolism, acetyl coenzyme a carboxylase 1, controls T cell immunity. J Immunol. 2014;192:3190–3199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 257.Berod L, Friedrich C, Nandan A, et al. De novo fatty acid synthesis controls the fate between regulatory T and T helper 17 cells. Nat Med. 2014;20:1327–1333. [DOI] [PubMed] [Google Scholar]
  • 258.Young KE, Flaherty S, Woodman KM, Sharma-Walia N, Reynolds JM. Fatty acid synthase regulates the pathogenicity of Th17 cells. J Leukoc Biol. 2017;102:1229–1235. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 259.McDonald G, Deepak S, Miguel L, et al. Normalizing glycosphingolipids restores function in CD4+ T cells from lupus patients. J Clin Invest. 2014;124:712–724. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 260.Sawaf M, Fauny JD, Felten R, et al. Defective BTLA functionality is rescued by restoring lipid metabolism in lupus CD4+ T cells. JCI Insight. 2018;3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 261.Thiam AR, Farese RV Jr., Walther TC. The biophysics and cell biology of lipid droplets. Nat Rev Mol Cell Biol. 2013;14:775–786. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 262.Shen Y, Wen Z, Li Y, et al. Metabolic control of the scaffold protein TKS5 in tissue-invasive, proinflammatory T cells. Nat Immunol. 2017;18:1025–1034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 263.Pearce EL, Poffenberger MC, Chang CH, Jones RG. Fueling immunity: insights into metabolism and lymphocyte function. Science. 2013;342:1242454. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 264.Weyand CM, Goronzy JJ. Immunometabolism in early and late stages of rheumatoid arthritis. Nat Rev Rheumatol. 2017;13:291–301. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 265.Chang CH, Curtis JD, Maggi LB Jr., et al. Posttranscriptional control of T cell effector function by aerobic glycolysis. Cell. 2013;153:1239–1251. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 266.Murphy MP, O’Neill LAJ. Krebs cycle reimagined: The emerging roles of succinate and itaconate as signal transducers. Cell. 2018;174:780–784. [DOI] [PubMed] [Google Scholar]
  • 267.Mills EL, Kelly B, O’Neill LAJ. Mitochondria are the powerhouses of immunity. Nat Immunol. 2017;18:488–498. [DOI] [PubMed] [Google Scholar]
  • 268.Brown PM, Pratt AG, Isaacs JD. Mechanism of action of methotrexate in rheumatoid arthritis, and the search for biomarkers. Nat Rev Rheumatol. 2016;12:731–742. [DOI] [PubMed] [Google Scholar]

RESOURCES